THE DYNAMICS OF COMPLEX BOX MAPPINGS

TREVOR CLARK, KOSTIANTYN DRACH, OLEG KOZLOVSKI, AND SEBASTIAN VAN STRIEN

Abstract. In holomorphic dynamics, complex box mappings arise as first return maps to well-chosen domains. They are a generalization of polynomial-like mapping, where the domain of the return map can have infinitely many components. They turned out to be extremely useful in tackling diverse problems. The purpose of this paper is: - To illustrate some pathologies that can occur when a complex box mapping is not induced by a globally defined map and when its domain has infinitely many components, and to give conditions to avoid these issues. - To show that once one has a box mapping for a rational map, these conditions can be assumed to hold in a very natural setting. Thus we call such complex box mappings dy- namically natural. Having such box mappings is the first step in tackling many problems in one-dimensional dynamics. - Many results in holomorphic dynamics rely on an interplay between combinatorial and analytic techniques. In this setting some of these tools are - the Enhanced Nest (a nest of puzzle pieces around critical points) from [KSS1]; - the Covering Lemma (which controls the moduli of pullbacks of annuli) from [KL1]; - the QC-Criterion and the Spreading Principle from [KSS1]. The purpose of this paper is to make these tools more accessible so that they can be used as a ‘black box’, so one does not have to redo the proofs in new settings. - To give an intuitive, but also rather detailed, outline of the proof from [KvS, KSS1] of the following results for non-renormalizable dynamically natural complex box mappings: - puzzle pieces shrink to points, - (under some assumptions) topologically conjugate non-renormalizable polynomials and box mappings are quasiconformally conjugate. - We prove the fundamental ergodic properties for dynamically natural box mappings. This leads to some necessary conditions for when such a box mapping supports a measurable invariant line field on its filled . These mappings are the analogues of Latt`es maps in this setting. - We prove a version of Ma˜n´e’sTheorem for complex box mappings concerning expansion along orbits of points that avoid a neighborhood of the set of critical points. arXiv:2105.08654v1 [math.DS] 18 May 2021

Date: May 19, 2021. Acknowledgments. The research of the second author was partially supported by the advanced grant 695 621 “HOLO- GRAM” of the European Research Council (ERC), which is gratefully acknowledged. We would also like to thank Dzmitry Dudko and Dierk Schleicher for many stimulating discussions and encouragement during our work on this project, Weixiao Shen and Mikhail Hlushchanka for helpful comments. 1 THE DYNAMICS OF COMPLEX BOX MAPPINGS 2

Contents 1. Introduction2 2. Examples of box mappings induced by analytic mappings 13 3. Examples of possible pathologies of general box mappings 21 4. Dynamically natural box mappings 26 5. Combinatorics and renormalization of box mappings 32 6. Statement of the QC-Rigidity Theorem and outline of the rest of the paper 37 7. Complex bounds, the Enhanced Nest and the Covering Lemma 43 8. The Spreading Principle and the QC-Criterion 53 9. Sketch of the proof of the QC-Rigidity Theorem 59 10. Sketch of the proof of Complex Bounds 62 11. All puzzle pieces shrink in diameter 67 12. Ergodicity properties and invariant line fields 67 13. Latt`esbox mappings 73 14. A Ma˜n´eTheorem for complex box mappings 79 References 84

1. Introduction

f V

Figure 1. The first return map to V .

In dynamical systems, a natural and effective way to understand the detailed behavior of a given map f is to consider the first return map under f to a certain set V , see Figure1. In this way, if the orbit of a point under iteration of f intersects V infinitely many times, then the first return map will send each point of intersection to the next point of intersection along the orbit. Therefore, by iterating the first return map instead of f, one can study THE DYNAMICS OF COMPLEX BOX MAPPINGS 3 properties of certain orbits of f “at a faster speed”. However, there is a trade-off: the first return map might have a rather complicated, even undesirable,1 structure. In the analytic setting, a return map might have the structure of a polynomial-like map. A polynomial-like map is a holomorphic branched covering F : U → V of degree at least two between a pair of open topological disks U ⊂ V such that U is relatively compact in V . The restriction of a complex polynomial to a sufficiently large topological disk in C is an example of a polynomial-like map. Such mappings are an indispensable tool in the field of holomorphic dynamics due to their fundamental role in renormalization and self-similarity phenomena. However, in general, the topological dynamics of a return mapping even under a polyno- mial cannot be described by a polynomial-like mapping. This motivates and explains the need for a more flexible class of mappings, namely, complex box mappings (see Figure2): Definition 1.1 (Complex box mapping). A holomorphic map F : U → V between two open sets U ⊂ V ⊂ C is a complex box mapping if the following holds: (1) F has finitely many critical points; (2) V is the union of finitely many open Jordan disks with disjoint closures; (3) every component V of V is either a component of U, or V ∩ U is a union of Jordan disks with pairwise disjoint closures, each of which is compactly contained in V ; (4) for every component U of U the image F (U) is a component of V, and the restriction F : U → F (U) is a proper map2.

Figure 2. An example of a complex box mapping F : U → V. The com- ponents of U are shaded in grey, there might be infinitely many of those; the components of V are the topological disks bounded by orange curves. The critical points of F are marked with crosses.

1For example, a component of the domain might not properly map onto a component of the range. 2In [KvS], where the definition of complex box mapping was given, the requirement that F maps each component of U onto a component of V properly was implicit. THE DYNAMICS OF COMPLEX BOX MAPPINGS 4

In the above definition of complex box mapping we assumed that each component of U and V is a Jordan domain. In some settings it is convenient to relax this, and assume that each component U of U and V of V a) is simply connected; b) when U b V , then V \ U is a topological annulus; c) has a locally connected boundary. In Section 6.2 we will discuss why and when the theorems in this paper will go through in this setting. In one-dimensional holomorphic dynamics, it is often the case that the first step in understanding the dynamics of a family of mappings is to obtain a good combinatorial model for the dynamics in the family. Once that is at hand, one can go on to study deeper properties of the dynamics in the family, for example, rigidity, ergodic properties and the geometric properties of the Julia sets. It turns out that complex box mappings are often part of these combinatorial models. Indeed, recent progress on the rigidity question for a large family of non-polynomial rational maps, so-called Newton maps [DS] made it clear that complex box mappings can be effectively used in the holomorphic setting well past polynomials: by “boxing away” the most essential part of the ambient dynamics into a box mapping one can readily apply the existing rigidity results without redoing much of the theory from scratch. These developments motivated us to give a comprehensive survey of the dynamics of such mappings. The subtlety is that Definition 1.1 is extremely flexible and thus allows for undesirable pathologies: Pathologies of General Complex Box Mappings. There exists • a complex box mapping with empty Julia set, • a complex box mapping with a set of positive measure that does not accumulate on the postcritical set, and • a complex box mapping with a wandering domain. These examples are given in Section3. We go on to introduce a class of complex box mappings, called dynamically natural (see Definition 4.1), for which none of these pathologies can occur. This class of mappings includes many of the complex box mappings induced by rational maps and real-analytic maps (and even in some weaker sense by a broad class of C3 interval maps), see Section2. Moreover, quite often general complex box mappings induce dynamically natural complex box mappings, see Proposition 4.5. We go on to give a detailed outline of the proof of qc (quasiconformal) rigidity of dy- namically natural complex box mappings. This result was first proved in [KvS] relying on techniques introduced in [KSS1]: QC Rigidity for Complex Box Mappings. Every two combinatorially equivalent non- renormalizable dynamically natural complex box mappings (satisfying some standard nec- essary conditions) are quasiconformally conjugate. We state a precise version of this result in Theorem 6.1. For the definition of combina- torial equivalence of box mappings see Definition 5.4. A dynamically natural complex box THE DYNAMICS OF COMPLEX BOX MAPPINGS 5 mapping is non-renormalizable if none of its iterates admit a polynomial-like restriction with connected filled Julia set; renormalization of box mappings is discussed in detail in Section 5.3. Since the proofs in [KvS] and [KSS1] are quite involved, and part of [KSS1] only deals with real polynomials, we will give a quite detailed outline of the proof of quasiconformal rigidity for complex box mappings (our Theorem 6.1). We will pay particular attention to the places in [KvS] where the assumption that the complex box mappings are dynamically natural is used while applying the results from [KSS1, Sections 5-6]. The main technical result that underlies qc rigidity are the complex or a priori bounds. These results give compactness properties for certain return mappings and control on the geometry of their domains. Complex bounds for non-renormalizable complex box mappings. Suppose that F : U → V is a non-renormalizable dynamically natural complex box mapping. Then there exist arbitrarily small combinatorially defined critical puzzle pieces P (components −n of F (V) for some n ∈ N) with complex bounds. See Theorem 7.1 for a precise statement of what we mean by complex bounds. Roughly speaking, for a combinatorially defined sequence Vn of pullbacks of V, complex bounds express how well return domains to Vn are inside of Vn in terms of a lower bound of the modulus of the corresponding annulus. One of the first applications of complex bounds in one-dimensional dynamics was in the study of the renormalization of interval mappings. Suppose that f is unimodal mapping with critical point at 0. One says that f is infinitely renormalizable if there exist a sequence {Ii} of intervals about 0 and an increasing sequence {si}, si ∈ N, so that si is the first s return time of 0 to Ii and f i (∂Ii) ⊂ ∂Ii. For certain analytic infinitely renormalizable mappings with bounded combinatorics, Sullivan [Su] proved that for all i sufficiently large, s we have that the first return mapping f i : Ii → Ii extends to a polynomial-like mapping Fi : Ui → Vi (see also [dMvS]). Moreover, he proved that one has the following complex bounds: there exist a constant δ > 0 and i0 so that mod(Vi \ U i) > δ for all i > i0. In fact, here δ is universal in the sense that it can be chosen so that it only depends on the order of the critical point of f, but i0 does depend on f. This property is known as beau bounds3. If there are several critical points we say that such a δ is beau if it depends only on the number and degrees of the critical points of f (and not on f itself). In fact, the complex bounds given by Theorem 7.1 for non-renormalizable box mappings are also beau for persistently recurrent critical points, see Section 5.2 for the definition. For reluctantly recurrent or non-recurrent critical points the estimates are not beau - they depend on the initial geometry of the box mapping, and there is no general mechanism for them to improve at small scales. We will not go into the history of results on complex bounds both for real and complex maps, but refer for references to [CvST]. That paper deals with real polynomial maps and shows that beau complex bounds hold in both the finitely renormalizable and the

3Beau, from French beautiful, nice, is a mixed acronym “a priori bounds that are eventually universal”. THE DYNAMICS OF COMPLEX BOX MAPPINGS 6 infinitely renormalizable cases, regardless whether critical points have even or odd order, or are persistently or not. This result is proved in [CvST]. In fact, in [CvST] these complex bounds are also established for real analytic maps (and even more general maps). Note that in the infinitely renormalizable case, such bounds cannot be obtained from Theorem 7.1. Indeed, in that case puzzle pieces do not shrink to points, and so one has to restart the initial puzzle partition at deeper and deeper levels. It turns out that for non-real infinitely renormalizable polynomials such complex bounds do not hold in general [Mil4]. As discussed below, complex bounds have many applications: that high renormalizations of infinitely renormalizable maps belong to a compact family of polynomial-like maps; local connectivity of the Julia sets; absence of measurable invariant line fields; quasiconformal rigidity for real polynomial mappings with real critical points; density of hyperbolicity in real one-dimensional dynamics, etc. In addition to their use in proving quasiconformal rigidity, we use complex bounds to study ergodic properties of complex box mappings. For example, we prove: Number of ergodic components. If F : U → V is a non-renormalizable dynamically natural complex box mapping, then for each ergodic component E of F there exist one or more critical points c of F so that c is a Lebesgue density point of E. In particular, the number of ergodic components of F is bounded above by the number of critical points of F . The analogous result in the real case was proved in [SV]. For rational maps there is the theorem of [Man] which states that each forward invariant set of positive Lebesgue measure accumulates to the ω-limit set of a recurrent critical point. The above result strengthens this in the setting of non-renormalizable dynamically natural complex box mappings. The above result is proved in Corollary 12.2 of Theorem 12.1, where we prove some fundamental ergodic properties of non-renormalizable dynamically natural complex box mappings. Our study of the ergodic properties of such box mappings leads us to some necessary conditions for when such a box mapping supports an invariant line field on its filled Julia set, see Proposition 13.2. Complex box mappings with a measurable invariant line field we call Latt`esbox mappings. Their properties are analogous to those of Latt`es rational maps. Latt`esbox mappings cannot arise for complex box mappings induced by polynomials, real maps or Newton maps; however, we give an example of a dynamically natural complex box mapping that is Latt`es,see Proposition 13.1. The rigidity theorem stated above deals only with non-renormalizable complex box map- pings. In the setting of dynamically natural mappings, if a map is non-renormalizable, then all its periodic points are repelling, see Section 5.3. However, it is possible, and in fact quite useful to work with complex box mappings that do admit attracting or neutral periodic cycles. Some results in this direction were obtained in [DS, Section 3]. In this paper, we push it a bit further and establish in Section 14 a Ma˜n´e-type theorem for complex box mappings. For non-renormalizable maps this result reads as follows: Ma˜n´e-type theorem for box mappings. Let F : U → V be a non-renormalizable dy- namically natural complex box mapping so that each domain is δ-nice (see Section 1.5.5). For each neighborhood B of Crit(F ) and each κ > 0 there exist λ > 1 and C > 0 so that THE DYNAMICS OF COMPLEX BOX MAPPINGS 7

k−1 k for all k > 0 and each x so that x, . . . , F (x) ∈ U \ B and d(F (x), ∂V) > κ, one has k k | DF (x)| > Cλ . In fact, we prove a slightly more general version of this theorem that does not depend on F being non-renormalizable, see Theorem 14.1. Also, this version of the theorem does not require that the orbits of points are bounded away from ω(c) for c ∈ Crit(F ) as in the classical Ma˜n´eTheorem for rational maps, see [Man]. When a holomorphic mapping induces a complex box mapping, such expansivity results are useful in the study of the measurable dynamics of the mapping and the geometry of their (filled) Julia sets. For example, this result can be applied to rational maps for which there is an induced complex box mapping that contains the orbits of all recurrent critical points intersecting the Julia set, see [Dr].

1.1. Some history of the notion. The idea of considering successive first return maps to neighborhoods of a critical point of an interval map is extremely natural, and was used extensively to show absence of wandering intervals and to obtain various metric properties. The notion of ‘box mapping’ was implicitly used in papers such as [BL,GJ, Jak,JS, NvS] and the terminology ‘nice interval’ (i.e. an interval V so that f n(∂V ) ∩ int(V ) = ∅ for n ∈ N) was introduced in Martens’ PhD thesis [Mar]. A natural way of obtaining such n −j intervals is by taking intervals in the complement of ∪j=0f (p), where p is a periodic point and n > 1. For complex mappings, box mappings are of course already implicit in Julia and Fatou’s work, after all that is how you can show that the Julia set of a quadratic map with the escaping critical point forms a : in this case, one sees that the filled Julia set of the polynomial is the filled Julia set of a complex box mapping with no critical point and whose domain has exactly two components. More generally, Douady and Hubbard went further by introducing the notion of a polynomial-like mapping where U and V each consist of just one component. Using this notion they were able to explain some of the similarities that one sees between the Julia sets of different mappings, similarities within the , and the appearance of sets which look like the Mandelbrot set in various families of mappings [DH1]. One beautiful property of polynomial-like mappings is the Douady–Hubbard Straightening Theorem: a polynomial-like mapping F is hybrid conjugate to a polynomial on a neighborhood of its filled Julia set; that is, there exist a polynomial P , a neighborhood U 0 of K(F ) and a quasiconformal mapping H so that P ◦ H(z) = H ◦ F (z) for z ∈ U 0 and ∂H¯ vanishes on K(F ). When the Julia set of F is connected, P is unique. Thus from a topological point of view, the family of polynomial-like mappings is no richer than the family of polynomials. Suppose one has a unicritical map F with a critical point c so that some iterate F s is maps some domain U 3 c as a branched covering onto V c U. If F (c) ∈ U for all s i > 0 then F is called renormalizable and F : U → V is a polynomial-like map. When F : U → V is not renormalizable, then there exists an element in the forward orbit of c that intersects the annulus V\U. In order to keep track of the full forward orbit of critical points, one would need to include in the domain of F all components of the domain of THE DYNAMICS OF COMPLEX BOX MAPPINGS 8 the return mapping which intersect the postcritical set. This is a typical use of a complex box mapping. If there are at most a finite number of such components, then the analogy between complex box mappings and polynomial-like mappings is strongest. Such mappings are also called polynomial-like box mappings or generalized polynomial-like mappings, and this notion played an important role in [LM, Lyu2]. However, in general, one needs to consider complex box mappings whose domains have infinitely many components, and even allow for several components in V. This motivates Definition 1.1. By allowing U to have infinitely many components this notion becomes very flexible at the expense of admitting pathological behavior, which we will discuss in Section3. In the literature, complex box mappings are sometimes called puzzle mappings, or R-mappings (where ‘R’ stands for return; see, for example, [ALdM]). Yoccoz gave a practical way of constructing complex box mappings induced by polyno- mials. Using the property that periodic points have external rays landing on them, Yoccoz introduced what are now known as Yoccoz puzzles [Hub1, Mil1]. In this case V consists of disks whose boundary consists of pieces of external rays and pieces of equipotentials. Yoc- coz used these puzzles to prove local connectivity of the Julia sets of non-renormalizable 2 quadratic polynomials, Pc : z 7→ z + c, and by transferring the bounds that he used to prove local connectivity of J(Pc) to the parameter plane, he proved local connectivity of the Mandelbrot set at parameters c for which Pc is non-renormalizable [Hub1]. One very important step in Yoccoz’ result is to obtain lower bounds for the moduli of certain annuli which one encounters when taking successive first return maps to puzzle pieces containing the critical point (the principal nest). Such a property is usually called complex bounds or a priori bounds. Yoccoz was able to obtain these bounds for non- renormalizable polynomial maps with a unique quadratic critical point that is recurrent: here one uses that the first return map to a puzzle piece has at least two components intersecting the critical orbit (this is related to the notion of children that we will discuss later on). It was clear from the early 1990s that complex methods would have wide applicability in one-dimensional dynamics, even when considering interval maps. In the 1990s complex bounds were established in the setting of real unimodal mappings, for certain infinitely renormalizable mappings in [Su], and for unimodal polynomials in [LvS1]. These results depended on additional properties of interval mappings. Soon further results were estab- lished for the quadratic family z 7→ z2 + c [Lyu2], and for classes of real-analytic mappings with a single critical point [McM, McM2, LY, LvS2]. Such results led to fantastic progress for real quadratic and unicritical maps:

• local connectivity of Julia sets [LvS1, Lyu2, LY, LvS2]; • monotonicity of entropy [MTh, Do2, Ts]; • density of hyperbolicity in the real quadratic family [Lyu2,GS] (via quasiconformal rigidity) • hyperbolicity of renormalization and the Palis Conjecture for mappings with a non- degenerate critical point [Lyu3, Lyu4, Lyu5, ALdM]. THE DYNAMICS OF COMPLEX BOX MAPPINGS 9

While some results held for unimodal maps with a higher degree critical point e.g. [Su, LvS1, McM, McM2], in general, it took some time for the theory for interval mappings with several critical points (or a critical point with degree greater than two) and for non-real polynomials, to catch up to that of quadratic mappings. In the setting of interval mappings, building on the work of [Su] (described in [dMvS]) and applying results of [McM2], complex bounds were obtained for certain infinitely renormalizable multicritical mappings together with a proof of exponential convergence of renormalization for such mappings [Sm1, Sm2]. Towards the goal of proving density of hyperbolic mappings for one-dimensional mappings, [Ko1] proved density of hyperbolicity for smooth unimodal maps and [She2] proved C2 density of Axiom A interval mappings by first proving complex bounds and certain local rigidity properties for certain multimodal interval mappings. To prove quasiconformal rigidity, the techniques from [Lyu2,GS] rely on the map to be unimodal and the critical point to be quadratic, because that implies that the moduli of certain annuli grow exponentially. To deal with general real multimodal maps, [KSS1] introduced a sophisticated combinatorial construction (the Enhanced Nest). Using this, together with additional results for interval maps, complex bounds and quasiconformal rigidity for real polynomial mappings with real critical points were proved in [KSS1]. This implies density of hyperbolicity for such polynomials. In [KSS2] complex box mappings in the non-minimal were constructed, and these were used to study ‘global perturbations’ of real analytic maps thus establishing density of hyperbolicity for one-dimensional mappings in the Ck topology for k = 1, 2,..., ∞, ω. One can also show that one has density of hyperbolicity for real transcendental maps and within full families of real analytic maps [RvS, CvS2]. One can apply the techniques of [KSS1] to prove complex bounds and qc rigidity for real analytic mappings and even in some sense for smooth interval maps, see [CvST, CvS]. The rigidity result from [KSS1] can also be used to extend the monotonicity of entropy for cubic interval maps [MTr] to general multimodal interval maps (with only real critical points) [BvS, Ko2] and to show that zero entropy maps or on the boundary of chaos [CT]. These results rely on complex bounds which were initially only available for real maps. Complex mappings lack the order structure of the real line, and some tools that are very useful for real mappings such as real bounds [SV] and Poincar´edisks cannot be used to prove complex bounds for complex mappings. Thus, new analytic tools were required, namely the Quasiadditivity Law and the Covering Lemma of [KL1]. Using this new ingredient, the theory for non-renormalizable complex polynomials was brought up to the same level as that of real quadratic polynomials in the unicritical setting in [KL2, AKLS, ALS] and in the general non-renormalizable polynomial case in [KvS]. Going beyond mappings with non-degenerate critical point, for real analytic unimodal maps with critical points of even degree, exponential convergence of renormalization was proved in [AL2] and the Palis Conjecture in this setting was proved in [Cl], see also [BSS]. The Covering Lemma was also used in [KvS] to prove local connectivity of Julia sets, absence of measurable invariant line fields and qc rigidity of non-renormalizable polynomials. THE DYNAMICS OF COMPLEX BOX MAPPINGS 10

1.2. The Julia set and filled Julia set of a box mapping. Given a complex box mapping F : U → V, the filled Julia set of F is defined as \ K(F ) := F −n(V). n>0 The set K(F ) consists of non-escaping points: those whose orbits under F never escape the domain of the map. As for the Julia set of F , there are a few candidates which one could take as the definition. No matter which definition that one uses, the properties of the Julia set in the context of complex box mappings are often quite different from those of the Julia set of a rational mapping, mainly due to the fact that U can have infinitely many components. We discuss this in Section 3.1, where we define two sets

JU (F ) := ∂K(F ) ∩ U and JK (F ) := ∂K(F ) ∩ K(F ), each of which could play the role of the Julia set. 1.3. Puzzle pieces of box mappings. One of the most important properties of complex box mappings is that they provide a kind of “Markov partition” for the dynamics of F on its filled Julia set: the components of V and their iterative pullbacks are analogous to Yoccoz puzzle pieces for polynomials. Having such a topological structure is a starting point for the study of many questions in dynamics. So, following [KvS, KSS1], for a given −n n > 0, a connected component of F (V) is called a puzzle piece of depth n for F . It is easy to see that F properly maps a puzzle piece of depth n + 1 onto a puzzle piece of depth n, for every n > 0, and any two puzzle pieces are either nested, or disjoint. In the former case, the puzzle piece at larger depth is contained in the puzzle piece of smaller depth. Remark. In the literature, the puzzle pieces that are constructed for polynomials or ratio- nal maps are usually assumed to be closed. In the context of polynomial-like mappings, or generalized polynomial-like mappings, similar to ours, the puzzle pieces are often defined as open sets. 1.4. Structure of the paper. The paper is organized as follows. In the last subsection of the introduction, Section 1.5, we introduce some notation and terminology that will be used throughout the paper. We encourage the reader to consult this terminological subsection if some notion was used in some section of the paper but was not defined earlier in that section. In Section2, we give several examples of complex box mappings that we have alluded to above and which appear naturally in the study of real analytic mappings on the interval and rational maps on the Riemann sphere. The purpose of this section is to give additional motivation for Definition 1.1 and to convince the reader that in a great variety of setups the study of a dynamical system can be almost reduced to the study of an induced box mapping, and hence one can prove rigidity and ergodicity properties of various dynamical systems simply by “importing” the general results on box mappings discussed in this paper. We end Section2 by posing some open questions about box mappings. THE DYNAMICS OF COMPLEX BOX MAPPINGS 11

In Section3, we present some of the pathologies that can occur for a box mapping due to the fairly general definition of this object. The discussion in that section will lead to the definition of a dynamically natural box mapping in Section4 for which such undesirable behavior cannot happen. These box mappings arise naturally in the study of both real and complex one-dimensional maps, hence the name. In Section5, we recall the definitions and properties of some objects of a combinatorial nature that can be associated to a given box mapping (fibers, recurrence of critical orbits, etc.). We end that section with the definitions of renormalizable and combinatorially equivalent box mappings. In Section6, we state the main result on the rigidity and ergodic properties of non- renormalizable dynamically natural box mappings, Theorem 6.1. This result includes shrinking of puzzle pieces, a necessary condition for a complex box mapping to support an invariant line field on its Julia set, and quasiconformal rigidity of topologically conjugate complex box mappings, with a compatible “external structure”. This result was proven in [KvS] for box mappings induced by polynomials (which never support invariant line fields on their filled Julia sets), and the proof extends to our more general setting. We will elab- orate on its proof with a two-fold goal: on one hand, to emphasize on several details and assumptions that were not mentioned or were implicit in [KvS], and on the other hand, to make the underlying ideas and proof techniques better accessible to a wider audience beyond the experts. The rest of the paper, namely Sections7–12, is dedicated to reaching this goal. The structure of the reminder of the paper, starting from Section7, is outlined in Section 6.3, and we urge the reader to consult that subsection for details. In Section 12 we build on the description of the ergodic properties of complex box mappings to give conditions for the absence of measurable invariant line fields for box mappings. In Section 13 we give examples of box mappings which have invariant line fields and which are the analogue of rational Latt`esmaps in this setting. We also derive a Ma˜n´e Theorem showing that one has expansion for orbits of complex box mappings which stay away from critical points, see Section 14. These results were not explicit in the literature.

1.5. Notation and terminology. We refer the reader to some standard reference back- ground sources in one-dimensional real [dMvS] and complex [Mil2, Lyu7] dynamics.

1.5.1. Generalities. We let C denote the complex plane, Cb be the Riemann sphere, we write D = D1 for the open unit disk and Dr for the open disk of radius r centered at the origin. A topological disk is an open simply connected set in C.A Jordan disk is a topological disk whose topological boundary is a Jordan curve. An annulus is a doubly connected open set in C. The conformal modulus of an annulus A is denoted as mod(A). By a component we mean a connected component of a set. For a set B and a connected subset C ⊂ B or a point C ∈ B, we write CompC B for the connected component of B containing C. THE DYNAMICS OF COMPLEX BOX MAPPINGS 12

We let A and cl A stand for the closure of a set A in C (we will use both notations interchangeably, depending on typographical convenience); int A is the interior of the set A. A domain A is compactly contained in a domain B, denoted by A b B, if A ⊂ B. We let diam(A) be the Euclidean diameter of a set A ⊂ C. A topological disk P ⊂ C has η-bounded geometry at x ∈ P if P contains the open round disk of radius η · diam(P ) centered at x. We simply say that P has η-bounded geometry if it has η-bounded geometry with respect to some point in it. For a Lebesgue measurable set X ⊂ C, we write meas(X) for the Lebesgue measure of X.

1.5.2. Notation and terminology for general mappings. In this paper, we will restrict our attention to the dynamics of holomorphic mappings. For a given holomorphic map f, f n denotes the n-th iterate of the map. We let Crit(f) denote the set of critical points of f, and PC(f) stands for the union of forward orbits of Crit(f), i.e. n PC(f) = {f (c): c ∈ Crit(f), n > 0} . For a point z in the domain of f, we write orb(z) for the forward orbit of z under f, n i.e. orb(z) = {f (z): n > 0}. We also write ω(z) for the ω-limit set of orb(z), i.e. ω(z) = T {f n(z): n > k}. k>0 For a holomorphic map f and a set B in the domain of f, a line field on B is the assignment of a real line through each point z in a positive measure subset E ⊂ B so that the slope is a measurable function of z. A line field is invariant if f −1(E) = E and the differential Dzf transforms the line at z to the line at f(z). A set A ⊂ U is minimal for a map f : U → V , U ⊂ V if the orbit of every point z ∈ A is dense in A. Periodic orbits are examples of minimal sets.

1.5.3. Nice sets and return mappings. For a map f, an open set B in the domain of f is called wandering if f k(B) 6= f `(B) for every k 6= ` and B is not in the basin of a periodic k . The set B is called nice if f (∂B) ∩ int B = ∅ for all k > 1; it is strictly nice if k f (∂B) ∩ B = ∅ for all k > 1. Given a holomorphic map f : U → V and a nice open set B ⊂ U, define k ˆ L(B) := {z ∈ U : ∃k > 1, f (z) ∈ B}, R(B) := L(B) ∩ B, L(B) := L(B) ∪ B. The components of L(B), R(B) and Lˆ(B) are called called, respectively, entry, return and landing domains (to B under f). For a point z ∈ U, we will use the following short notation: ˆ ˆ Lz(B) := Compz L(B), Lz(B) := Compz L(B). k(z) The first entry map EB : L(B) → B is defined as z 7→ f (z), where k(z) is the minimal k(z) positive integer with f (z) ∈ B. The restriction of EB to B is the first return map to B; it is defined on R(B) and is denoted by RB. The first landing map LB : Lˆ(B) → B is defined as follows: LB(z) = z for z ∈ B, and LB(z) = EB(z) for z ∈ L(B) \ B. THE DYNAMICS OF COMPLEX BOX MAPPINGS 13

1.5.4. Notation and terminology for complex box mappings. In this paper, our object of study is a complex box mapping F : U → V. Sometimes will abbreviate this terminology and simply refer to a box mapping F , and unless otherwise stated, by box mapping we mean a complex box mapping. −n We define the puzzle pieces of depth n to be the components of F (V). If A ⊂ C, we call a collection of puzzle pieces whose union contains A a puzzle neighborhood of A. The non-escaping (or filled Julia) set of F is defined as K(F ) := T F −n(V). n>0 A box mapping F 0 : U 0 → V0 is induced from F : U → V if U 0 and V0 are unions of puzzle pieces of F and the branches of F 0 are the iterates of the branches of F . Given a point z ∈ K(F ) a nest of puzzle pieces about z or simply a nest is a sequence of puzzle pieces that contain z. We say that puzzle pieces shrink to points if for any infinite nest P1 ⊃ P2 ⊃ P3 ⊃ ... of puzzle pieces, we have that diam(Pn) tends to zero as n → ∞. We will say that a complex box mapping F : U → V has moduli bounds if there exists δ > 0 so that for every component U of U that is compactly contained in a component V of V we have that mod(V \ U) > δ.

1.5.5. δ-nice and δ-free puzzle pieces. A puzzle piece P is called δ-nice if for any point z ∈ PC(F ) ∩ P, whose orbit returns to P , we have that mod(P \ Lz(P )) > δ. Note that strictly nice puzzle pieces are automatically 0-nice. A puzzle piece P is called δ-free4, if there exist puzzle pieces P − and P + with P − ⊂ P ⊂ P +, so that mod(P + \ P ) and mod(P \ P −) are both at least δ and the annulus P + \ P − does not contain the points in PC(F ).

2. Examples of box mappings induced by analytic mappings In this section, we give quite a few examples of complex box mappings that appear in the study of complex rational and real-analytic mappings. We end this section with an outlook of further perspectives and open questions.

2.1. Finitely renormalizable polynomials with connected Julia set, see Figure3. Suppose that P : C → C is a polynomial mapping of degree d with connected Julia set. d By B¨ottcher’s Theorem, P restricted to C \ K(P ) is conjugate to z 7→ z on C \ D by a conformal mapping ϕ: C\D → C\K(P ). There are two foliations that are invariant under z 7→ zd: the foliation by straight rays through the origin, and the foliation by round circles centered at the origin. Pushing these foliations forward by ϕ, one obtains two foliations of C \ K(P ), which are invariant under P . The images of the round circles under ϕ are called equipotentials for P , and the images of the rays are called rays for P . Quite often the terminology dynamic rays is used to distinguish these rays from rays in parameter space, but we will not need to make that distinction in this paper. When constructing the Yoccoz puzzle, an important consideration is whether the rays land on the filled Julia set (that is, the limit set of the ray on the filled Julia set consists of a single point). Fortunately, rays

4In [KvS], the terminology δ-fat was used instead of δ-free. In the present paper, we prefer the latter, more friendly and more transparent terminology. THE DYNAMICS OF COMPLEX BOX MAPPINGS 14 always land at repelling periodic points and consequently, such rays form the basis for the construction of the Yoccoz puzzle.

Figure 3. The Yoccoz puzzle pieces of depths 0, 1 and 2 for the polynomial z 7→ z2 + i are shown in color on the left, middle and right picture respec- tively. In this example, Y0 consists of three pieces shown in green, blue and orange on the left. The pieces at lager depths are iterative preimages of pieces at smaller depths and are shown in corresponding colors.

To construct the Yoccoz puzzle, one needs to be able to make an initial choice of an equipotential and rays which land at repelling periodic or preperiodic points and separate the plane5. The top level Y0 of the Yoccoz puzzle is then defined as the union of bounded components of C in the complement of the chosen equipotential and rays, see Figure3. For −n 0 n > 0, a Yoccoz puzzle piece of depth n is then a component of P (Y ). For unicritical mappings, z 7→ zd + c, to define Y0, one typically takes the rays landing at the dividing fixed points (fixed points α so that K(P ) \{α} is not connected), their preimages, and an arbitrary equipotential. The main feature of Yoccoz puzzle pieces is that they are nice, i.e. P n(∂V ) ∩ V = ∅ for every piece V and n ∈ N. This property guarantees that puzzle pieces at larger depths are contained in puzzle pieces at smaller depths (compare Figure3). Moreover, if all the rays in the boundary of the puzzle piece V land at strictly preperiodic points, then V is strictly n nice: P (∂V )∩V = ∅ for n ∈ N. In this case, all of the return domains to V are compactly contained in V , and so the first return mapping to V has the structure of a complex box mapping. This construction is the prototypical example of a complex box mapping. In general, if P is at most finitely renormalizable, that is P has at most finitely many distinct polynomial-like restrictions, then as was shown in [KvS, Section 2.2], for such polynomials with additional property of having no neutral periodic points one can always find a strictly nice neighborhood V of the critical set of P , with this neighborhood being

5This can be done, for example, when all periodic points of the polynomial are repelling. THE DYNAMICS OF COMPLEX BOX MAPPINGS 15 a finite union of puzzle pieces, so that the first return mapping to V under P has the structure of a complex box mapping.

2.2. Real analytic mappings, see Figure4. For a real analytic mapping f of a compact interval, in general, one does not have Yoccoz puzzle pieces. Recall that to construct the Yoccoz puzzle for a polynomial P one makes use of the conjugacy between P and zd in a neighborhood of ∞, where d is the degree of P . Nevertheless, one can construct by hand complex box mappings, which extend the real first return mappings, in the following sense. There exists a nice neighborhood I ⊂ R of the critical set of f with the property that each component of I contains exactly one critical point of f, and a complex box mapping F : U → V, so that the real trace of V is I, and for any component U of U, the real trace of U is a component of the (real) first return mapping to I. Suppose that f is an analytic unimodal mapping with critical point c. If there exists a nice interval I 3 c, so that so-called scaling factor |I|/|Lc(I)| is sufficiently big, then one can construct a complex box mapping which extends the real first return mapping to Lc(I) by taking V to be a Poincar´elens domain with real trace Lc(I), i.e. see Figure4. It turns out that for non-renormalizable unimodal mappings with critical point of degree two or a reluctantly recurrent critical point of any even degree (see Section 5.2 for the definition), one can always find such a nice neighborhood I of c so that |I|/|Lc(I)| is arbitrarily large. For infinitely renormalizable mappings, and for non-renormalizable mappings with degree greater than 2, this need not be the case. If all the scaling factors are bounded, then the construction is trickier, see for example [LvS2]. Such complex box mappings were constructed for general real-analytic interval mappings in [CvST, CvS], see also [Ko1] for unicritical real analytic maps with a quadratic critical point.

V U

c Figure 4. An example of a complex box mapping F : U → V associated to a real-analytic unimodal mapping f with critical point c whose orbit is dense in the interval. V is a Poincar´elens domain such that V ∩ R is a nice interval containing c, and U is the domain of the first return map to V under f. In this example, U ∩ R is dense in V ∩ R, i.e. the real trace of U “tiles” the real trace of V. THE DYNAMICS OF COMPLEX BOX MAPPINGS 16

2.3. Newton Maps, see Figure5. Given a complex polynomial P : C → C, the Newton map of P is the rational map NP : Cb → Cb on the Riemann sphere Cb defined as P (z) N (z) := z − . P P 0(z) These maps are coming from Newton’s iterative root-finding method in numerical analysis and hence provide examples of well-motivated dynamical systems. The roots of P are attracting or super-attracting fixed points of P , while the only remaining fixed point of NP in Cb is ∞ and it is repelling. The set of points in Cb converging to a root is called the basin of this root, and the component of the basin containing the root is the immediate basin. Newton maps are arguably the largest family of rational maps, beyond polynomials, for which several satisfactory global results are known. This progress was possible due to abun- dance of touching points between the boundaries of components of the root basins: these touchings provide a rigid combinatorial structure. Similarly to the real-analytic mappings, Newton maps do not have global B¨ottcher coordinates. However, the local coordinates in each of the immediate basins of roots allow one to find local equipotentials and local (internal) rays that nicely co-land from the global point of view. In this way, Newton maps posses forward invariant graphs (called Newton graphs) that provide a partition of the Riemann sphere into pieces similar to Yoccoz puzzle pieces, see Figure5. Contrary to the Yoccoz construction, building the Newton puzzle is not a straightforward task. For Newton maps of degree 3 it was carried out in [Roe], while for arbitrary degrees in was done in a series of papers [DMRS, DLSS], and independently in [WYZ]. Once the Newton puzzles are constructed, one can induce a complex box mapping as the first return map to a certain nice union of Newton puzzle pieces containing the critical set; this was done in [DS], where the rigidity results for box mappings that we discuss in the present paper were applied to conclude rigidity of Newton dynamics.

2.4. Other examples. Let us move to further examples of dynamical systems existing in the literature where a complex box mapping with nice dynamical properties can be induced.

2.4.1. Box mappings from nice couples. In [R-L], the concept of nice couples is defined. Given a rational map f : Cb → Cb, a nice couple for f is a pair of nice sets (V,Vˆ ) such that V b Vb, each component of V,Vb is an open topological disk that contains precisely one n element of Crit(f)∩J(f), and so that for every n > 1 one has f (∂V )∩Vb = ∅. Note that if (V,Vb ) is a nice couple, then V is strictly nice. This implies that if V := V and F : U → V is the first return map to V under f, then each component of U is compactly contained in V. It then follows that F is a complex box mapping in the sense of Definition 1.1. We see that if f has a nice couple, then f induces a complex box mapping F : U → V such that U is the return domain to V under f and Crit(f) ⊂ V. In [PRL1, PRL2], the authors study thermodynamic formalism for rational maps that have arbitrarily small nice couples. They show that certain weakly expanding maps, like topological Collet–Eckmann rational mappings (which include rational mappings that are THE DYNAMICS OF COMPLEX BOX MAPPINGS 17

Figure 5. An example of the Newton puzzle partition for the degree 4 Newton map. The dynamical plane of the map is shown on the left, with the roots of the corresponding polynomial being marked with circles. For each n > 0, the Newton puzzle partition of depth n is a tiling of a neighbor- hood of the Julia set on the Riemann sphere with topological disks (Newton puzzle pieces). These partitions have nice mapping properties (the Markov property), and as n grows, they become finer and finer: the number of pieces of depth n grows and depth n + 1 pieces tile the pieces of depth n (modulo truncation in the basins of roots). On the right, the Newton puzzle partition of depth 0 is shown (7 puzzle pieces are drawn in different colors). exponentially expanding along their critical orbits), have nice couples, and hence induce complex box mappings (see also [RLS]).

2.4.2. Box mappings associated to Fatou components. In several scenarios, one can con- struct a complex box mapping associated to a periodic Fatou component of a rational map. One instance when it is particularity easy to do is when a rational map f possess a fully invariant infinitely connected attracting Fatou component U. An example of such a rational map is a complex polynomial with an escaping critical point. For such f, as a starting set of puzzle pieces one can use, for instance, the disks bounded by the level lines of a Green’s function associated to U for some sufficiently small potential (see Figure6). The first return mapping under f to the union of those disks that contain the non-escaping critical points would then induce a complex box mapping. The results from the current paper, namely, those described in Theorem 6.1, Sections 6.1 and 12.1, can be then used to deduce various rigidity and ergodicity results for such f’s. Thus one can rephrase (and possibly shorten) the proofs in [QY], [YZ] and [Zh], where the authors deal with the above mentioned class of rational maps, by using the language and machinery of complex box mappings more explicitly. THE DYNAMICS OF COMPLEX BOX MAPPINGS 18

2 : 1

1 : 1

−1 1 U2

U1 V

Figure 6. The filled Julia set of the cubic polynomial P : z 7→ c−(z3 −3z− 2)/(c2 − c − 2) for c = −0.2 + 0.1i. In this example, Crit(P ) = {−1, 1} with 1 an escaping critical point and −1 of period two. Since the critical point escapes, the complex box mapping F : U1 t U2 → V can be induced using −1 the equipotentials ∂V and ∂U1 ∪ ∂U2 = P (∂V) in the basin of ∞. Puzzle pieces of larger depth are shown in deeper shades of red. This example illustrates a standard way of inducing a box mapping for rational maps with an infinitely connected fully invariant attracting Fatou component.

In [RY], it was shown that for a complex polynomial P every bounded Fatou component U, which is not a Siegel disk, is a Jordan disk. For the proof, it is enough to consider the situation when U is an immediate basin of a super-attracting fixed point (in the parabolic case, the corresponding parabolic tools should be used instead, see [PR]). The authors then build puzzle pieces in a neighborhood of ∂U by using pairs of periodic internal (w.r.t. U) and external rays that co-land on ∂U, as well as equipotentials in U and in the basin of ∞ for P . Using these puzzle pieces, it is possible to construct a strictly nice puzzle neighborhood of Crit(P ) ∩ ∂U. The first return map to this neighborhood then defines a complex box mapping. Using this box mapping, the main result of [RY] follows from Theorem 6.1 and the discussion in Section 6.1. A similar strategy was used in [DS] in order to prove local connectivity for the boundaries of root basins for Newton maps. 2.4.3. Box mappings in McMullen’s family. In [QWY], puzzle pieces were constructed for n n certain McMullen maps fλ : z 7→ z + λ/z , λ ∈ C \{0}, n > 3. This family includes maps with a Sierpi´nskicarpet Julia set. In contrast to the previously discussed examples, where the puzzle pieces where constructed using internal and external rays, the pieces constructed THE DYNAMICS OF COMPLEX BOX MAPPINGS 19 in [QWY] are bounded by so-called periodic cut rays: these are forward invariant curves that intersect the Julia set in uncountably many points and whose union separates the plane. Properly truncated in neighborhoods of 0 and ∞ (the latter is a super-attracting fixed point of fλ, and fλ(0) = ∞), these curves and their pullbacks provide an increasingly fine subdivision of a neighborhood of J(fλ) into puzzle pieces. Using these pieces, a complex box mapping can be induced. Similarly to the examples above, the main results of [QWY] can be then obtained by importing the corresponding results on box mappings.

2.5. Outlook and questions. In this subsection we want to mention some research ques- tions related to the notion of complex box mapping and to the techniques presented in this paper.

2.5.1. Combinatorial classification of analytic mappings via box mappings. The classifica- tion of box mappings goes via some combinatorial construction hinted at in Figure8. For polynomials and rational maps one often uses trees and tableaux to obtain combinatorial information. This is encoded in the pictograph introduced in [dMP] for the case where infinity is a super attracting fixed point whose basin is infinitely connected. It would be interesting to explore the relationship in more depth.

2.5.2. Metric properties of analytic mappings via box mappings. Complex box mappings play a crucial role in the study of the measure theoretic dynamics of rational mappings and the fractal geometry of their Julia sets. This has been a very active area of research, and here we just provide a snapshot of some of the results in these directions. Some of the natural questions in this setting concern the following: (1) The existence and properties of conformal measures supported on the Julia set, and the existence and properties of invariant measures that are absolutely continuous with respect to a conformal measure. (2) Finding combinatorial or geometric conditions on a mapping that have conse- quences for the measure or Hausdorff dimension of its Julia set, i.e. if the measure is positive or zero, or whether the Hausdorff dimension is two or less than two. (3) When is the Julia set holomorphically removable? (4) There are several different quantities which are related to the complexity of a fractal or expansion properties of a mapping on its Julia set, among them, the Hausdorff dimension, the hyperbolic dimension, and the Poincar´eexponent. While it is known that these quantities are not always all equal [AL3], in many circumstances they are, and it would be very interesting to characterize those mappings for which equality holds. The most complete results are known for rational mappings that are weakly expanding on their Julia sets, for example see [PRL2, RLS]. Both conformal measures and absolutely continuous invariant measures as in (1) are well-understood. It is known that whenever the Julia set of such a mapping is not the whole sphere that it has Hausdorff dimension less than two, and for such mappings all the aforementioned quantities in (4) are equal. Moreover, when additionally such a mapping is polynomial, its Julia set is removable. THE DYNAMICS OF COMPLEX BOX MAPPINGS 20

For infinitely renormalizable mappings with bounded geometry and bounded combina- torics [AL1] establishes existence of conformal measures; equality of the quantities men- tioned in (4), when the Julia has measure zero; and that for such mappings if the Hausdorff dimension is not equal to the hyperbolic dimension, then the Julia set has positive area. The existence of mappings with positive area (hence with Hausdorff dimension two), but hyperbolic dimension less than two was proved in [AL3]. A final class of mappings for which many such results are known are non-renormalizable quadratic polynomials. It was proved in [Lyu1, Shi] that their Julia sets have measure zero, in [Ka] that they are removable, and in [Prz] that their Hausdorff and hyperbolic dimensions coincide. Further progress on these questions for rational maps (or polynomials) will most likely involve complex box mappings, and moreover any such questions could also be asked for complex box mappings. Indeed this point of view is taken in for example [PZ]. 2.5.3. When can box mappings be induced? Once one has a box mapping for a given analytic map, one can use the tools discussed in this paper. Therefore it is very interesting to find more classes of maps for which box mappings exist: Question 1. Is it true that for every rational (or meromorphic) map with a non-empty Fatou set one has an associated non-trivial box mapping? Here we say that a box mapping is non-trivial if the critical set of the box mapping is non-empty and it satisfies the no permutation condition (see Definition 4.1). In fact, the authors are not aware of any general procedure which associates a non-trivial box mapping to a transcendental function. Examples of box mappings in the complement of the postsingular set for transcendental maps were constructed in [Dob]. Question 2. Are there examples of rational maps whose Julia set is the entire sphere and for which one cannot find an associated non-trivial box mapping? For example, for topological Collet–Eckmann rational maps, [PRL2] uses the strong expansion properties to construct box mappings. The previous two questions ask whether one can also do this when there is no such expansion. Similarly: Question 3. Can one associate a box mapping to a rational map with a Sierpi´nskicarpet Julia set beyond the examples discussed previously (where symmetries are used)? Another class of maps for which the existence of box mappings is not clear is when one has neutral periodic points: Question 4. Let f be a complex polynomial with a Siegel disk S. Under which conditions is it possible to construct a puzzle partition of a neighborhood of ∂S and using this partition to induce a complex box mapping? There is a recent result in this direction, namely, in [Ya] a puzzle partition was con- structed for polynomial Siegel disks of bounded type rotation number. Using this par- tition and the Kahn–Lyubich Covering Lemma (see Lemma 7.10), local connectivity of THE DYNAMICS OF COMPLEX BOX MAPPINGS 21 the boundary of such Siegel disks was established. In that paper, the author exploits the Douady–Ghys surgery and the Blaschke model for Siegel disks with bounded type rotation number. This model allows one to construct “bubble rays” growing out of the boundary of the disk. These rays, properly truncated, then define the puzzle partition. However, the resulting puzzle pieces have more complicated mapping properties than traditional Yoccoz puzzles (for example, they develop slits under forward iteration). Hence it is not clear at this point whether the tools presented in this paper can be applied (see also Section 6.2).

3. Examples of possible pathologies of general box mappings The goal of this section is to point out some “pathological issues” that can occur if we consider a general box mapping, without knowing that it comes from a (more) globally defined holomorphic map. We start with the following result: Theorem 3.1 (Possible pathologies of general box mappings). There are complex box mappings Fi : Ui → Vi, i ∈ {1, 2, 3} with the following properties:

(1) We have that K(F1) = V1 and JU (F1) = ∅. (2) The filled Julia set K(F2) has full Lebesgue measure in U2, empty interior, and there exists a positive (indeed full) measure set of points in K(F2) that does not accumulate on any critical point. Moreover, both JK (F2) and K(F2) carry invariant line fields. (3) V3 is a disk and each connected component U of U3 is compactly contained in V3 and contains a wandering disk for F3.

These examples are constructed so that Ui, Vi and Fi are symmetric with respect to the real line. Remark. Assertion (3) shows that for general box mappings the diameter of an infinite sequence of distinct puzzle pieces Pk does not need to shrink to zero as k → ∞, even if their depths tend to infinity. Assertion (2) shows that, even though a complex box mapping may be “expanding”, its Julia set can have positive measure.

Proof. To prove (1), take U1 = V1 = D and F1 is the identity map. Then K(F1) = V1 and JU (F1) = ∅ (in particular K(F1) is not closed). The example of (2) is based on the Sierpi´nskicarpet construction. Consider the square V2 := (−1, 1) × (−1, 1) cut into 9 congruent sub-squares in a regular 3-by-3 grid, and let U1 be the central open sub-square. The same procedure is then applied recursively to the remaining 8 sub-squares; this defines U2 as the union of the 8 central open sub-sub-squares. Repeating this ad infinitum, we define an open set U2 to be the union of all Ui. Note that U2 has full Lebesgue measure in V2 because the Lebesgue measure of V2 \ (∪i6nUi) is equal to 4(8/9)n. Define F2 on each component U of U2 as the affine conformal surjection from U onto V . Since K(F ) = T K , where K := F −n(V ), the set K(F ) also has full Lebesgue 2 2 n>1 n n 2 2 2 measure in U2, and JK (F2) = K(F2) as K(F2) has no interior points. Clearly, the horizontal line field in both K(F2) and JK (F2) is invariant under F2. THE DYNAMICS OF COMPLEX BOX MAPPINGS 22

To prove (3), take a monotone sequence of numbers ai ∈ (0, 1) such that ai % 1 and ∞ Y (3.1) ai = 1/2. i=1

Construct real M¨obiusmaps gi : D → D inductively as follows. Let g1 be the identity map. Take g2 such that it maps D onto D and Da2 to some disk to the right of g1(Da1 ) = Da1 . Then assuming that g1, . . . , gk−1 are defined for some k > 2, define gk to be so that it maps D onto D and so that gk(Dak ) is strictly to the right of gk−1(Dak−1 ). It follows that gk(Dak ) is disjoint from gi(Dai ) for all i < k. Next define a box mapping F : U → V (which will play the role of F3 : U3 → V3) by taking

[ −1 V = D1, U = gk(Dak ),F (x) = gk+1(gk (x)/ak) if x ∈ gk(Dak ) k>1 (see Figure7).

Figure 7. The wandering disk construction in Theorem 3.1(3): the colored disks are components of the domain of the box mapping F , while the gray shaded disks are part of the trajectory of the wandering disk W . THE DYNAMICS OF COMPLEX BOX MAPPINGS 23

Let us show that W := D1/2 is a wandering disk. Observe that (3.1) implies a1 ·...·an > 1/2 for every n > 1, and thus a1 · ... · an > |x| for every x ∈ W . Therefore, if x ∈ W , then −1 x ∈ g1(W ) and so F (x) = g2(g1 (x)/a1) = g2(x/a1) ∈ g2(Da2 ). Similarly, 2 −1 F (x) = g3(g2 (g2(x/a1))/a2) = g3(x/(a1a2)) ∈ g3(Da3 ). n Continuing in this way, for each x ∈ W and each n > 0 we have F (x) = gn+1(x/(a1 · ... · n an)). It follows that F (W ) ⊂ gn+1(Dan+1 ) and therefore W is a wandering disk. 

3.1. A remark on the definitions of K(F ),JU (F ) and JK (F ). There is no canoni- cal definition of the Julia set of a complex box mapping, so we have given two possible contenders: JU (F ) = ∂K(F ) ∩ U and JK (F ) = ∂K(F ) ∩ K(F ). In routine examples, neither JK (F ) nor K(F ) is closed, but JU (F ) is relatively closed in U. Moreover, JU (F ) can strictly contain K(F ). While the definitions of K(F ),JU (F ) and JK (F ) are similar to the definitions of the filled Julia set and Julia set of a polynomial-like mapping, when a complex box mapping has infinitely many components in its domain, the properties of its K(F ),JU (F ) and JK (F ) can be quite different. For example, let F : U → V be a complex box mapping associated to a unimodal, real-analytic mapping f : [0, 1] → [0, 1], with critical point c and the property that its critical orbit is dense in [f(c), f 2(c)] (see Figure4 and the discussion about real- analytic maps in Section 2.2 on how to construct such a box mapping). Then V will be a small topological disk containing the critical point, and U, the domain of the return mapping to V will be a union of countably many topological disks contained in V with the property that U ∩ R is dense in V ∩ R. In this case, one can show that [ −n K(F ) ∩ R = (V ∩ R) \ F (E), n>0 where E is the hyperbolic set of points in the interval whose forward orbits under f avoid V. Thus K(F ) ∩ R is a dense set of points in the interval V ∩ R. Thus JU (F ) is the union of open intervals U ∩ R, and it is neither forward invariant nor contained in the filled Julia set. Nevertheless it is desirable to consider JU (F ), since it agrees with the set of points in U at which the iterates of F do not form a normal family. 3.2. An example of a box mapping for which a full measure set of points con- verges to the boundary. In this section, we complement example (2) in Theorem 3.1 by showing that not only we can have the non-escaping set of a general box mapping F : U → V to be of full measure in U, but also almost all points in K(F ) are “lost in the boundary” as their orbits converges to the boundary under iteration of F ; an example with such a pathological behavior is constructed in Proposition 3.2 below. Note that the box mapping F2 : U2 → V2 constructed in Theorem 3.1(2) had no critical points. It is not hard to modify this example so that the modified map has a non-escaping critical point. Indeed, let Vˆ2 := V t U for some topological disk U, Uˆ2 := U2 t U, and ˆ ˆ ˆ ˆ ˆ ˆ define a map F2 : U2 → V2 by setting F2|U2 = F2, F2(U) = V2 and so that F2|U is a branched covering of degree at least two so that the image of a critical point lands in K(F2). This critical point for Fˆ2 will be then non-escaping and non-recurrent. Contrary to THE DYNAMICS OF COMPLEX BOX MAPPINGS 24 this straightforward modification, the box mapping constructed in the proposition below has a recurrent critical point and the construction is more intricate. Proposition 3.2 (Full measure converge to a point in the boundary). There exists a complex box mapping F : U → V with Crit(F ) ∩ K(F ) 6= ∅ and with the property that the set of points z ∈ K(F ) whose orbits converge to a boundary point of V has full measure in U. Proof. Let V the square with corners at (3/2, 3/2), (−3/2, 3/2), (−3/2, −3/2) and (3/2, −3/2). We will construct U so that it tiles V. Let S0 be the open square with side-length one cen- tered at the origin. It has corners at (1/2, 1/2), (−1/2, 1/2), (−1/2, −1/2), and (1/2, −1/2). Let S1 be the union of twelve open squares, Qi,j, each with side-length 1/2, surrounding S0, so that S1 together with S0 tiles the square with corners at (1, 1), (−1, 1), (−1, 1), and (1, −1). We call S1 the first shell. Inductively, we construct the i-th shell as the union of i open squares with side length 1/2 , surrounding Si−1, so that S0 ∪ S1 ∪ · · · ∪ Si−1 ∪ Si tiles i the square centered at the origin with side length 2(3/2 − 1/2 ). Inside of each square Qi,j in shell i, we repeat the Sierpi´nksicarpet construction of Theorem 3.1(2). These open sets, which consist of small open squares, together with the central component S0, will be the domain of the complex box mapping we are constructing. Let us fix a uniformization ψ : D → Qˆ, where Qˆ is the square with side length 3, which we may identify with V by translation. Fix any i ∈ N, and let R be a square given by the Sierpi´nskicarpet construction in the i-th shell. Let Ai be the linear mapping that rescales R so that Ai(R) has side length 3. We will define F |R : R → V by −1 F |R = ψ ◦ Mi ◦ ψ ◦ Ai, where Mi is a M¨obiustransformation that we pick inductively. To make the construction explicit, we take Mi(−1) = −1, Mi(1) = 1 and determine Mi by choosing a point zi ∈ 0 (−1, 0) so that Mi(0) = zi. Note that for any discs D,D centered at −1 respectively 1 one 0 can choose zi so that Mi(D \ D) ⊂ D ∩ D . For later use, let K be so that for any univalent map ϕ: U → V and for any square Qi,j from the initial partition the following inequality holds: | Dϕ(z)| K | Dϕ(z0)| 6 for all z, z0 ∈ U with ϕ(z), ϕ(z0) ∈ R. j For any j ∈ N ∪ {0}, we let Tj = ∪k=0Sk ⊂ D. To construct M1, we choose z1 so close −1 to ∂D that the set of points in R which are mapped to T1 by F |R = ψ ◦ M1 ◦ ψ ◦ A1 has area at most meas(R)/2. Up to different choices of rescaling, we define F in the same way on each component R contained in S1. Assuming that Mi−1, i ≥ 2, has been chosen, let R be a square in Si and pick zi ∈ (−1, 0) close enough to −1 so that meas({z ∈ R : F (z) ∈ L }) 1 (3.2) i > 1 − , meas(R) K2 · 2i where −i −i −i Li := (−3/2, −3/2 + 2 ) × (−2 , 2 ). THE DYNAMICS OF COMPLEX BOX MAPPINGS 25

Again, up to rescaling, we define F identically on each component of the domain in Si. Continuing in this way, we extend F to each shell. k Let z0 ∈ Si0 , and for k ∈ N define ik so that F (z0) ∈ Sik . We say that the orbit of z0 escapes monotonically to ∂V if ik is a strictly increasing sequence. Let W denote the set ∞ of points in ∪i=1Si whose orbits escape monotonically to (−3/2, 0) ∈ ∂V .

Claim. There exists C0 > 0 so that for any i > 1 and any component of the domain R in Si, meas(W ∩ R)/meas(R) is bounded from below by C0.

Proof of the claim. Note that the square Li is disjoint from Ti and that Li is a union of squares from the initial partition used to define the shells. Hence W ∩ R contains the set 2 of points so that {z ∈ R; F (z) ∈ Li,F (z) ∈ Li+1,..., }. To obtain a lower bound for the Lebesgue measure of this set, let R(F k(z)) be the rectangle containing F k(z) and notice that by (3.2) and Koebe 2 k k+1 meas({z ∈ R; F (z) ∈ Li,F (z) ∈ Li+1,...,F (z) ∈ Lk,F (z) ∈ Lk+1}) −i 2 k > 1 − 2 . meas({z ∈ R; F (z) ∈ Li,F (z) ∈ Li+1,...,F (z) ∈ Lk)} The claim follows. X 2 Let us now define F on S0. Let f : z 7→ z , Y be the disk of radius 16 centered at the −1 origin and X = f (Y ). Choose real symmetric conformal mappings H0 : S0 → X and z−a H1 : V → Y which send the origin to itself. Let ma = 1−az¯ be the family of real symmetric M¨obiustransformations of Y . Consider the family of mappings −1 Fa = H1 ◦ ma ◦ f ◦ H0 : S0 → V.

For a = 0,Fa is a real-symmetric, polynomial-like mapping with a super-attracting critical point at 0 of degree 2, which is a minimum for the mapping restricted to its real trace. As a varies along the positive real axis from 0, the critical value of the mapping Fa varies along the negative real axis from 0 until it escapes V\ S0. Thus the family of mappings

Fa|S0 is a full real family of mappings [dMvS], and hence it contains a mapping conjugate to the real Fibonacci mapping. Let F : S0 → V denote this mapping. Now we have defined F : U → V, where U tiles V. Let us now show that under F a full measure set of points in U converge to (−3/2, 0) ∈ ∂V . First, recall that the filled Julia set K(f) of the quadratic Fibonacci mapping has measure zero [Lyu1]. Since F : S0 → V is a polynomial-like mapping that is quasicon- formally conjugate to the quadratic Fibonacci mapping, and quasiconformal mappings are absolutely continuous, we have that the filled Julia set of F : S0 → V has measure zero. We will denote this set by K0. Since the set of points whose orbits eventually enter K0 is contained in the union of the countably many preimages of S0, we have that the set of points that eventually enter K0 has measure zero too. From the construction of F , we have that every puzzle piece contains points that map to S0, so we have that almost every point in K(F ) either accumulates on the critical point of F or converges to ∂V . By the comment above, we may assume that this set of points is disjoint from the preimages of K0. THE DYNAMICS OF COMPLEX BOX MAPPINGS 26

i Let us show that a.e. point in X1 = {z; F (z) ∈/ S0, i > 0} converges to (−3/2, 0). 0 Suppose not. Then there exists a set X1 ⊂ X1 for which this is not the case. Let z0 be a 0 Lebesgue density point of X1. Let Qk be the puzzle piece containing z0 of level k. Then k−1 F (Qk) = R for some rectangle R with R ∩ S0 = ∅. By the above claim and the Koebe Distortion Theorem a definite proportion of Qk is mapped into the set W of points which converge monotonically to (−3/2, 0), thus contradicting that z0 is a Lebesgue density point 0 of X1. Let X0 be the set of points which enter S0 and which are not eventually mapped into the 0 zero Lebesgue measure set K0. Let X0 be the set of points in X0 which do not converge 0 to (−3/2, 0) and let z0 be a Lebesgue density point of X0. Note that by the previous 0 paragraph a.e. point in X0 enters S0 infinitely many times. Let r0 ≥ 0 be minimal so that r k F 0 (z0) is contained in S0. Let k0 > r0 be minimal so that F 0 (z0) ∈ Si for some i > 0, and k0 let Rk0 denote the square of Si that contains F (z0). Inductively, define ri > ri−1 minimal ri ki so that F (z0) ∈ S0, and ki > ri, minimal so that F (z0) ∈ Si, i > 0, and let Rki be the ki square so that F (z0) ∈ Rki . By the choice of K at the start of the proof, the distortion ki −ki of the mapping F |Qi : Qi → Rki where Qi = Compz0 F (Rki ) is bounded by K. We have already proved that in each rectangle R ⊂ Si, i > 1 a definite proportion of points converge monotonically to the boundary point (−3/2, 0); hence at arbitrarily small scales around z0, a definite mass of points escapes is mapped into the set W (which converge to 0 (−3/2, 0) ∈ ∂V), which implies that z0 cannot be a Lebesgue density point of X0. 

4. Dynamically natural box mappings In this section we introduce the concept of a dynamically natural complex box mapping. These are the maps for which various pathologies from Section3 disappear, and which arise naturally in the study of rational maps on Cb. In order to define the concept, let us start by introducing two dynamically defined subsets of the non-escaping set.

4.1. Orbits that avoid critical neighborhoods. The first subset consists of points whose orbits avoid a neighborhood of Crit(F ). Let A ⊂ K(F ) be a finite set and W be a union of finitely many puzzle pieces. We say that W is a puzzle neighborhood of A if A ⊂ W and each component of W intersects the set A. If Crit(F ) 6= ∅, define

Koff-crit(F ) := {x ∈ K(F ): ∃W puzzle neighborhood of Crit(F ) : orb(x) ∩ W = ∅} ; otherwise, i.e. when F has no critical points, we set Koff-crit(F ) ≡ K(F ). It is easy to see that the set Koff-crit(F ) is forward invariant with respect to F .

4.2. Orbits that are well inside. The second subset consists of any point in K(F ) whose orbit from time to time visits components of U that are well-inside of the corresponding components of V. More precisely, let  mF (x) := mod Compx V\ Compx U , THE DYNAMICS OF COMPLEX BOX MAPPINGS 27 and for a given δ > 0, set   k Kδ(F ) := y ∈ K(F ) : lim sup mF (F (y)) > δ . k>0 Define [ Kwell-inside(F ) := Kδ(F ). δ>0

A component U of U is said to be δ-well-inside V if mF (x) > δ for some (and hence for all) x ∈ U. Thus the set Kwell-inside(F ) consists of the points x ∈ K(F ) whose orbit visits infinitely often components of U that are δ-well inside for some δ = δ(x) > 0. By definition, the set Kwell-inside(F ) is forward invariant with respect to F .

4.3. Dynamically natural box mappings. Definition 4.1 (No permutation condition and dynamical naturality). A complex box mapping F : U → V satisfies the no permutation condition if n (1) for each component U of U there exists n > 0 so that F (U) \ U 6= ∅. F : U → V is called dynamically natural if it satisfies additionally the following assumptions:

(2) the Lebesgue measure of the set Koff-crit(F ) is zero; (3) K(F ) = Kwell-inside(F ). If F does not satisfy the no permutation condition, then there exists a component U of U and an integer k > 0 so that F k(U) = U. Notice that this implies that U, . . . , F k−1(U) are all components of V and that F cyclically permutes these components. Let us return to the pathologies described in Theorem 3.1. Each of the box mappings Fi, i ∈ {1, 2, 3}, is not dynamically natural in the sense of Definition 4.1: for the map Fi the respective condition (i) in that definition is violated. We see that

• the box mapping F1 has no escaping points in U1; • the non-escaping set of F2 is equal to Koff-crit(F2), and hence F2 provides an example of a box mapping with Koff-crit(F2) of non-zero Lebesgue measure; • the wandering disk constructed for F3 does not belong to Kwell-inside(F3), and hence K(F3) \ Kwell-inside(F3) is non-empty. Moreover, the example constructed in Proposition 3.2 is also not a dynamically natural box mapping as it violates condition (2) of the definition of naturality. The following lemma implies that no dynamically natural box mapping has the pathology described in Theorem 3.1(1). Lemma 4.2 (Absence of components with no escaping points). If F : U → V is a complex box mapping that satisfies the no permutation condition (Definition 4.1), then

• JK (F ) 6= ∅; • each component of K(F ) is compact; • if U has finitely many components, then K(F ) and JU (F ) = JK (F ) are compact. THE DYNAMICS OF COMPLEX BOX MAPPINGS 28

Proof. Take a nested sequence P1 ⊃ P2 ⊃ ... of puzzle pieces. Then JK (F ) 6= ∅ because for each i > 0 either Pi+1 = Pi or Pi+1 is compactly contained in Pi. If F satisfies the no T permutation condition (condition (1) of Definition 4.1), then necessarily Pi is compactly contained in V and so JK (F ) 6= ∅. The second and third assertions also follow. 

4.4. Motivating the notion of ‘dynamically natural’ from Definition 4.1. First of all, condition (1) prohibits F to simply permute components of U. Equivalently, it guarantees that each component of U has escaping points under iteration of F . This is clearly something one should expect from a mapping induced by rational maps: for example, under this simple condition, as we saw in Lemma 4.2, each component of K(F ) is a compact set. Furthermore, under this condition we can further motivate assumption (2), see the re- mark after Corollary 4.4: for such complex box mappings a.e. point either converges to the boundary of V or accumulates to the set of critical fibers. For any known complex box mapping which is induced by a rational map, the boundary of V does not attract a set of positive Lebesgue measure, and therefore automatically (2) holds. Thus it makes sense to assume (2). Assumption (3) is imposed because one is usually only interested in points that visit the ‘bounded part of the dynamics’ infinitely often. In fact, as we will show in Proposition 4.5, under the no permutation condition it is always possible to improve any box mapping to be dynamically natural by taking some further first return maps; this way of “fixing” general box mappings is enough in many applications.

Remark. In [ALdM] a different condition was used to rule out the pathologies we discussed in Section3. That paper is concerned with unicritical complex box mappings F : U → V, where V consists of a single domain. In [ALdM] it is assumed that U is thin in V where U is called thin in V if there exist L,  > 0, such that for any point z ∈ U there is a open topological disk D ⊂ V of z with L-bounded geometry at z such that mod(V\ D) > , and meas(D\U)/ meas(D) > . See Section 5 and Appendix A of [ALdM] for results concerning this class of mappings. In particular, when U is thin in V we have that Koff-crit(F ) has measure zero. On the other hand, the requirement that U is thin is V is a stronger geometric requirement than K(F ) = Kwell-inside(F ) (compare Figure 15).

4.5. An ergodicity property of box mappings. In this subsection, we study an ergod- icity property in the sense of typical behavior of orbits of complex box mappings that are not necessarily dynamically natural, but satisfy condition the no permutation condition from Definition 4.1. The results of this subsection will be used later in Section 4.6, where we will show how to induce a dynamically natural box mapping starting from an arbitrary one, and in Section 12, we will strengthen the results in this subsection and study invariant line fields of box mappings. Let us say that F is a critical fiber of F if it is the intersection of all the puzzle pieces containing a critical point. We say that this set is a recurrent critical fiber if there exist THE DYNAMICS OF COMPLEX BOX MAPPINGS 29

n iterates ni → ∞ so that some (and therefore all) limit points of F i (F) are contained in F (we refer the reader to Section5 for a detailed discussion on fibers and types of recurrence). Lemma 4.3 (Ergodic property of general box mappings). Let F : U → V be a complex box mapping that satisfies the no permutation condition. Let C be the union of recurrent critical fibers of F . Define i X1 = {z ∈ K(F ): F (z) → ∂V as i → ∞},

ij X2 = {z ∈ K(F ): F (z) → C for some sequence ij → ∞} and X = K(F ) \ (X1 ∪ X2). Then X is forward invariant. Moreover, if X0 ⊆ X is a forward invariant set of positive Lebesgue measure, then there exists a puzzle piece J of F so that meas(J ∩X0) = meas(J ).

Proof. That X is forward invariant is obvious. For any integer n > 0, let Yn be the union of all critical puzzle pieces of depth n for F containing a recurrent critical fiber. Define En ⊂ K(F ) to be the set of points whose forward orbits are disjoint from Yn. By construction, (E ) is a growing sequence of forward invariant sets and X0 ⊂ S E . If X0 n n>0 n 0 has positive Lebesgue measure, then there exists n0 so that En0 ∩X has positive Lebesgue 0 measure. Let z0 ∈ En0 ∩ X be a Lebesgue density point of this set. 0 0 n Starting with V := V, U := U and F0 := F , for each n > 1 inductively define Fn : U → n n V to be the first return map under Fn−1 to the union V of all critical components of n−1 U . Note that Crit(Fn) ⊂ Crit(F ) and K(Fn) contains all recurrent critical fibers of F (it is straightforward to see that fibers of Fn are also fibers of F for each n > 0). We may assume that there exists k > 0 so that the Fk-orbit of z0 visits non-critical k components of U infinitely many times. Indeed, otherwise the F -orbit of z0 accumulates at a recurrent critical fiber contrary to the definition of En0 . Fix this mapping Fk. n k Case 1: the orbit zn := Fk (z0), n = 0, 1,... visits non-critical components of U infinitely many times, but only finitely many different ones. Write U 0 for the union of 0 n 0 these finitely many components and consider K := {z ∈ En0 : Fk (z) ∈ U ∀n > 0}. By 0 Lemma 4.2, the closure of K is a subset of En0 , and hence of K(F ). Therefore, there exists a point y ∈ ω(z0) ∩ En0 . Let J0 3 y be a puzzle piece of depth n0, and let D b J0 be another puzzle piece containing y of depth larger than n0; such a compactly contained ni puzzle piece exists again because of Lemma 4.2. Take the moments ni so that Fk (z0) ∈ J0 −ni ni and let Ai := Compz0 Fk (J0). Then Fk : Ai → J0 is a sequence of univalent maps −ni with uniformly bounded distortion when restricted to Bi := Compz0 Fk (D). Therefore mod(Ai \ Bi) is uniformly bounded from below. We also have that Ai0 is contained in Bi 0 for i > i sufficiently large. It follows that diam Bi → 0 as i → ∞. Thus, by the Lebesgue 0 0 Density Theorem, meas(D ∩ En0 ) = meas(D), and hence meas(D ∩ X ) = meas(D) as X is forward invariant. The claim of the lemma follows with J = D. Case 2: the orbit zn visits infinitely many different non-critical components of Fk. If k ns is the first moment the orbit zn visits a new non-critical component Us of U , then the pullbacks of Us under Fk back to z0 along the orbit hit each critical fiber at most once. THE DYNAMICS OF COMPLEX BOX MAPPINGS 30

0 Let ns > ns be the first subsequent time that the orbit zn enters a critical component 0 0 k 0 ns −ns k k of U . If such an integer ns exists, then we have that Fk : Compz0 Fk (V ) → V has degree bounded independently of s. We can proceed again as in Case 1 (there are k 0 only finitely many critical components of U ). If ns does not exist, then we have that n −n k k Fk : Compz0 Fk (V ) → V has uniformly bounded degree for all n > ns. Since we assumed that z0 does not converge to the boundary, there exists an accumulation point k k k y ∈ V ∩ω(z0) with the property that J0 := Compy V is not a component of U . Note that k k k y ∈ U , but y may or may not lie in U . Each component of U ∩J0 is compactly contained in J0. This allows us to choose an open topological disk D b J0 so that y ∈ D and such that D contains at least one puzzle piece J . We conclude similarly to the previous case 0 0 that meas(D ∩ X ) = meas(D). Thus meas(J ∩ X ) = meas(J ) as J ⊂ D.  Corollary 4.4 (Typical behavior of orbits). Let F : U → V be a complex box mapping that satisfies the no permutation condition and such that each puzzle piece of F either contains n a point w so that F (w) is a critical point for some n > 0, or it contains an open set disjoint from K(F ) (a gap). Then for a.e. z ∈ K(F ) either (1) F i(z) → ∂V as i → ∞, or (2) the forward orbit of z accumulates to the fiber of a critical point. 0 Proof. Let X ⊂ K(F ) be as in the previous lemma. For each n > 1, let Xn ⊂ X be the set of points z ∈ X so that the orbit of z remains outside critical puzzle pieces of depth n. By 0 the previous lemma if Xn has positive Lebesgue measure, then there exists a puzzle piece 0 J so that meas(J ∩ Xn) = meas(J ). But, under the assumption of the corollary, either J 0 contains a gap, which is clearly impossible since meas(J ∩ Xn) = meas(J ), or there exists 0 a puzzle piece Jn ⊂ J which is mapped under some iterate of F to a critical piece of depth 0 0 0 0 n, which contradicts the definition of the set Xn since also meas(Jn ∩ Xn) = meas(Jn). 0 0 0 Hence the set Xn has zero Lebesgue measure. It follows that the set X = ∪Xn of points in X whose forward orbits stay outside some critical puzzle piece also has measure zero. The corollary follows.  Remark. For a polynomial or rational map, preimages of critical points are dense in the Julia set (except if the map is of the form z 7→ zn or z 7→ z−n). This means that if a box mapping is associated to a polynomial or rational map, then the assumption of Corollary 4.4 is satisfied. The conclusion of that corollary motivates the assumption meas(Koff-crit(F )) = 0 in the definition of dynamical naturality (Definition 4.1). 4.6. Improving complex box mappings by inducing. We end this section by showing how to “fix” a general complex box mapping for it to become dynamically natural. In fact, every box mapping satisfying the no permutation condition from the definition of naturality, subject to some mild assumptions, induces a dynamically natural box mapping that captures all the interesting critical dynamics; working with such induced mappings is enough in many applications. The following lemma illustrates how it can be done in detail. Proposition 4.5 (Inducing dynamically natural box mapping). Let F : U → V be a box mapping that satisfies the no permutation condition. Assume that each component of V THE DYNAMICS OF COMPLEX BOX MAPPINGS 31 contains a critical point, or contains an open set which is not in U. Then we can induce a dynamically natural box mapping from F in the following sense. Take Uˆ to be a finite union of puzzle pieces of F so that Uˆ is a nice set compactly contained in V, and let CritUˆ(F ) ⊂ Crit(F ) be a subset of critical points whose orbits visit Uˆ infinitely many times. Then there exists an induced, dynamically natural box mapping

F1 : U1 → V1 such that CritUˆ(F ) ⊂ Crit(F1) ⊂ Crit(F ). Remark. A typical choice for Uˆ would be puzzle pieces of U that contain critical points and critical values of F .

0 00 Proof of Proposition 4.5. Let us define two disjoint nice sets V1, V1 ⊂ V as follows. For the first set, we let 0 [ ˆ ˆ V1 := Lc U, c∈Crit(F ) where the union is taken over all points in Crit(F ) whose orbits intersect Uˆ. For the second 00 set, we define V1 to be the union of puzzle pieces of depth n > 0 containing all critical 0 00 points in Crit(F ) \V1 (if this set is empty, we put V1 = ∅). Moreover, we can choose n 00 00 sufficiently large so that V1 ∩ V1 = ∅ and so that V1 is compactly contained in V; the latter can be achieved due to Lemma 4.2. 0 00 Now let V1 := V1 ∪ V1 , and let F1 : U1 → V1 be the first return map to V1 under F . It easy to see that F1 is a complex box mapping in the sense of Definition 1.1; it also satisfies property (1) of Definition 4.1 (notice that K(F1) ⊂ K(F )). Moreover, from the first return construction it follows that CritUˆ(F ) ⊂ Crit(F1) ⊂ Crit(F ). Let us now show that F1 is dynamically natural. For this we need to check properties (2) and (3) of Definition 4.1.

Property (2) of Definition 4.1: Let XO ⊂ Koff-crit(F1) be the set of points in K(F1) whose forward iterates under F1 avoid a puzzle neighborhood O of Crit(F1); this is a forward invariant set. Hence, if XO has positive Lebesgue measure, then by Lemma 4.3 either a set of positive Lebesgue measure of points in XO converge to the boundary of V1, or there exists a puzzle piece J of F1 so that meas(J ∩ XO) = meas(J ). The first situation ˆ 00 cannot arise due to the choice of U and V1 (both are compactly contained in V), and the second situation would imply that some forward iterate of J will cover a component of V1 and therefore a component V of V, and thus we would obtain meas(XO ∩ V ) = meas(V ), contradicting to the assumption we made on F : the puzzle piece V must contain either a critical point, or an open set in the complement of U.

Property (3) of Definition 4.1: If there are only finitely many components in U1, this property is obviously satisfied. Therefore, we can assume that U1 has infinitely many s components. Let U be a component of U1 such that U b CompU V1, and let F : U → V be the corresponding branch of F1; here V is a component of V1. By the first return construction, the degree of this branch is bounded independently of the choice of U. By construction of V1, there exists t > 0 and a component W of V1 such that W b V and F s+t(U) = W . Put s0 := s + t. Again, the degree of the map F s0 : U → W is bounded THE DYNAMICS OF COMPLEX BOX MAPPINGS 32

0 independently of U. Let W c W be a puzzle piece of F ; it exists since W b V. We claim s0 0 0 0 −s0 0 that the degree of the map F : U → W , where U = CompU F (W ), is independent of the choice of U. Indeed, this is equivalent of saying that in the sequence F (U 0),...,F s0−1(U 0) the number of critical puzzle pieces is independent of the choice of U. Let P be a critical puzzle piece in this sequence, and c ∈ Crit(F ) ∩ P be a critical point. Then P ) Compc V1, because otherwise the F -orbit of U would intersects V1 before reaching V , which is a contradiction s to the fact that F : U → V is the branch of the first return map to V1. But there are only finitely many critical puzzle pieces of F of depths larger than the depths of the components of V1, and this number does not depend on U. This yields the claim. Pulling back the annulus W 0\W by the map F s0 : U 0 → W 0, we conclude that there exists 0 0 δ > 0, independent of U, such that mod(U \ U) > δ. Finally, U ⊂ V1 because the depths 0 0 of U and U , viewed as puzzle piece of F , differ by one. Hence (CompU V1) \ U ⊃ U \ U. This together with the moduli bound imply K(F1) = Kδ(F1). 

5. Combinatorics and renormalization of box mappings The combinatorics of critical points of complex box mappings are similar to those of complex polynomials. As is often the case in holomorphic dynamics, the tools used to study the orbit of a critical point depend on the combinatorial type of the critical point. In this section, we start by recalling the definitions fibers, of recurrent, persistently, and reluctantly recurrent critical points, and of combinatorial equivalence of box mappings. We end this section by defining the notion of renormalization for complex box mappings. In this section, F : U → V is a complex box mapping, not necessarily dynamically natural, but that satisfies the no permutation condition from Definition 4.1.

5.1. Fibers. While working with dynamically defined partitions, for example given by puzzles, it is convenient to use the language of fibers, introduced by Schleicher in [Sch1, Sch2]. In our setting, for x ∈ K(F ), let Pn(x) be the puzzle piece of depth n containing x. The fiber 6 of x is the set \ fib(x) := Pn(x). n>0 By Lemma 4.2, fib(x) is a compact connected set. From the dynamical point of view, fib(x) consists of points that are “indistinguishable” by the puzzle partition and hence constitute a single combinatorial class. A fiber containing a critical point is called critical.

6The term “fiber” was adopted from the theory of Douady’s pinched disk models for quadratic Julia sets and the Mandelbrot set: in this context, there is a natural continuous surjection from a set to its pinched disk model, and the preimage of a point under this surjection — i.e. the fiber in the classical sense — is exactly a single combinatorial fiber in the sense of [Sch1, Sch2]. For details, see [Do1, Sch3]. THE DYNAMICS OF COMPLEX BOX MAPPINGS 33

5.2. Persistently and reluctantly recurrent critical points. A critical point c ∈ Crit(F ) is said to be (combinatorially) non-recurrent if the orbit of F (c) is disjoint from some puzzle piece around c. If this is not the case, then c is called (combinatorially) recurrent. For such points, orb(F (c)) intersects every puzzle neighborhood of c7. In the language of fibers, a critical point c is combinatorially recurrent if and only if the orbit of F (c) accumulates at fib(c). It will be useful to define Back(c) := {c0 ∈ Crit(F ): orb(c0) ∩ fib(c) 6= ∅}, Forw(c) := {c0 ∈ Crit(F ): orb(c) ∩ fib(c0) 6= ∅}, to be the sets of critical points “on the backward, resp. forward orbit of the critical point c”, and [c] := Back(c) ∩ Forw(c) to be the set of critical points that “accumulate at each other’s fibers”. Clearly, [c] 6= ∅ only if c is combinatorially recurrent. Note that Back(c) = Back(c0), Forw(c) = Forw(c0), and [c] = [c0] for every c0 ∈ fib(c). A critical puzzle piece P is a child of a critical puzzle piece Q if there exists n > 1 such that F n−1 : F (P ) → Q is a univalent map. Now we want to distinguish different type of recurrence as follows. Let c be a combi- natorially recurrent critical point of F . We say that c is persistently recurrent if for every 0 0 critical point c ∈ [c], every puzzle piece Pn(c ) has only finitely many children containing critical points in [c]. Otherwise, c is called reluctantly recurrent. One should observe that all critical points in a critical fiber are either simultaneously non-recurrent, or all persistently recurrent, or all reluctantly recurrent, and hence we can speak about non-, reluctantly, or persistently recurrent critical fibers8.

5.3. Renormalization. Following [KvS, Definition 1.3], a complex box mapping F : U → 9 V is called (box) renormalizable if there exists s > 1 and a puzzle piece W of some depth ks containing a critical point c ∈ Crit(F ) such that F (c) ∈ W for all k > 0. If this is not the case, then F is called non-renormalizable. Lemma 5.1 (Douady–Hubbard equivalent to box renormalization). If F : U → V is a com- plex box mapping that satisfies the no permutation condition, then F is box renormalizable if and only if it is renormalizable in the classical Douady–Hubbard sense [DH1].

7In [KSS1], the notion of Z-recurrent points appears. There, Z refers an initial choice of puzzle partition and recurrence is defined with respect to that initial choice. In the setting of a given complex box mapping, one works with a single puzzle partition that is predefined by the map. 8This observation demonstrates the idea that a given puzzle partition cannot distinguish points from their fibers. In this vein, following Douady (with an interpretation by Schleicher), various rigidity results can be phrased as “fibers are points”, i.e. each fiber is trivial and hence equal to a point. 9Usually we assume that s is minimal with this property. Such minimal s is called the period of renormalization. THE DYNAMICS OF COMPLEX BOX MAPPINGS 34

Recall that one says that a box mapping F : U → V is Douady–Hubbard renormalizable if there exist s ∈ N, c ∈ Crit(F ) and open topological disks c ∈ U b V ⊂ U so that F s : U → V is a polynomial-like map with connected Julia set. Proof of Lemma 5.1. First suppose that F is box renormalizable, and let c ∈ Crit(F ), W 3 c and s be as in the definition of box renormalizable. By Lemma 4.2, the fiber of c is ks a compact connected set. The condition F (c) ∈ W for all k > 0 can be now re-interpreted ks 0 0 as F (c ) ∈ fib(c) ⊂ W for all k > 0 and all critical points c ∈ fib(c). Now let n be large −ns enough so that the only critical points of F in the set V := Compc F (W ) are the one −s contained in the fiber of c. If U := Compc F (V ), then U b V because F satisfies the no permutation condition (see assumption (1) in Definition 4.1), and hence F s : U → V is a polynomial-like map. This mapping has connected Julia set equal to fib(c) by our choice of n. We conclude that F is renormalizable in the Douady–Hubbard sense. Conversely, suppose that F : U → V is Douady–Hubbard renormalizable. Then there s exist c ∈ Crit(F ), s ∈ N and a pair of open topological disks U b V so that % := F : U → V is a polynomial-like map with connected Julia set K(%) with c ∈ K(%). Note that U and V are not necessarily puzzle pieces of F , however K(%) ⊂ fib(c). The last inclusion implies that for every puzzle piece W of F with W 3 c it is also true that W ⊃ K(%). Since s ks K(%) is forward invariant under F , we have that for all k ∈ N, F (c) ∈ W . Picking a smaller W to assure minimality of s in the definition of box renormalization, we conclude the claim.  We now want to relate the existence of non-repelling periodic points for a complex box mapping to its renormalization. Lemma 5.2 (Non-repelling cycles imply renormalization). If F : U → V is a complex box mapping that satisfies the no permutation condition, then the orbit of the fiber of any attracting or neutral periodic point contains a critical point of F . Moreover, if F is non- renormalizable, then all the periodic points of F are repelling. 0 Proof. Suppose that z0 is a periodic point with period p for F . Let U0 denote the component −p 0 of U that contains z0, and let U0 = Compz0 F (U0). Since F satisfies the no permutation 0 condition (see assumption (1) in Definition 4.1), we have that U0 b U0. We claim that the p 0 mapping F : U0 → U0 has a critical point. If not, then consider the inverse of this map. By the Schwarz Lemma this inverse map will be strictly contracting at its fixed point.  5.4. Itineraries of puzzle pieces and combinatorial equivalence of box mappings. By Definition 1.1, each component in the sets U and V for a box mapping F : U → V is a Jordan disk. Therefore, by the Carath´eodory theorem for every branch F : U → V of this mapping there exists a well-defined continuous homeomorphic extension Fˆ : U → V , and by continuity this extension is unique. Let us denote by Fˆ : cl U → cl V the total extended map. Definition 5.3 (Itinerary of puzzle pieces relative to curve family). Let F : U → V be a box mapping satisfying the no permutation condition, and let X ⊂ ∂V be a finite set with one point on each component of ∂V. Let Γ be a collection of simple curves in (cl V)\(U∪PC(F )), THE DYNAMICS OF COMPLEX BOX MAPPINGS 35 one for each y ∈ Fˆ−1(X), that connects y to a point in X (see Figure8). Then for every −n n > 0 and for each component U of F (U) there exists a simple curve connecting X to k n ∂U of the form γ0 . . . γn where Fˆ (γk) ∈ Γ. The word (γ0, Fˆ(γ1),..., Fˆ (γn)) is called the Γ-itinerary of U. If a component V of V is also a component of U, then the corresponding curve will be contained in the Jordan curve (cl V )\(U ∪PC(F )) = ∂V , because by definition PC(F ) ⊂ V and so no point in the boundary of V lies in PC(F ). Note the Γ-itinerary of U is not uniquely defined even though there is a unique finite word for every y0 ∈ Fˆ−n(x)∩∂U. For example, in Figure8 the critical component of F −1(U) has four Γ-itineraries: (a, a), (a, b), (b, a), and (b, b). However, different components of F −n(U) have different Γ-itineraries. Definition 5.3 might look a bit unnatural at first. A simpler approach might to define an itinerary of a component U of F −n(U) as a sequence of components of U through which the F -orbit of U travels. However, such an itinerary would not distinguish components properly, an example is again shown in Figure8: both A1 and A2 have the same “itineraries through the components of U”, but clearly A1 6= A2. This motivates using curves in order to properly distinguish components. Remark (On the existence of Γ). We observe that a collection Γ in Definition 5.3 always exists. Indeed, let U, V be a pair of components of U resp. V with U ⊂ V , and x ∈ ∂V , y ∈ ∂U be a pair of points we want to connect with a curve within (cl V) \ (U ∪ PC(F )). Clearly, the set A := PC(F ) ∩ (V\ U) is finite (the box mapping F is not defined in V\ U). Choose any simple curve γ : [0, 1] → V \ U so that γ(0) = x and γ(1) = y. Such a curve exists because V \ U is a closed non-degenerate annulus bounded by two Jordan curves. We can choose γ so to avoid A as well as the closures of the components W of U in V with ∂W ∩ PC(F ) 6= ∅. The set A is finite, and there are only finitely many such W ’s. Hence a curve γ with such an additional property exists. This curve might intersect a component 0 0 U 6= U of U within V . Let t1 be the infimum over all t ∈ [0, 1] so that γ(t) ∈ ∂U ; similarly, 0 let t2 be the supremum over all such t’s (it follows that 0 < t1 6 t2 < 1 as ∂U ∩ ∂U = ∅ 0 0 and ∂U ∩ ∂V = ∅). Modify the curve γ on [t1, t2] by substituting it with the part of ∂U 0 going from γ(t1) to γ(t2). Since ∂U is a Jordan curve, the new curve (we keep calling it γ) will be a Jordan curve. Moreover, γ will avoid U 0, will join x to y and will be disjoint from A. The desired curve in Γ is obtained after resolving all intersections with all such U 0. Finally, let us give a definition of combinatorially equivalent box mappings (see [KvS, Definition 1.5]).

Definition 5.4 (Combinatorial equivalence of box mappings). Let F : U → V and Fe : Ue → Ve be two complex box mappings, both satisfying the no permutation condition. Let H : V → Ve be a homeomorphism with the property that H(U) = Ue, H(PC(F ) \U) = PC(Fe) \ Ue and such that it has a homeomorphic extension Hˆ : cl V → cl Ve to the closures of V and Ve. THE DYNAMICS OF COMPLEX BOX MAPPINGS 36

Figure 8. An example of a complex box mapping F : U → V with U = U1 ∪U2 (shaded in light grey), V having one component and a unique simple critical point (marked with the red cross) which mapped by F to U1. An example of a curve family Γ = {a, b, c} is shown. The set F −1(U) has 5 components (shown in dark grey). Two of these, marked A1 and A2, are mapped over U2 and hence have the same itineraries with respect to the components of U. However, their Γ-itineraries are different: A1 has Γ-itinerary (b, c), while the Γ-itinerary of A2 is (a, c).

The maps F and Fe are called combinatorially equivalent w.r.t. H if there exists a curve family Γ as in Definition 5.3 so that if Γe = Hˆ (Γ), then each critical point c ∈ Crit(F ) is mapped by H to a critical point ec ∈ Crit(Fe) with the property that for every integer −n k > 0 and n > 0 the Γ-itineraries of CompF k(c) F (V) coincide with the Γ-itinerariese of −n Comp k Fe (Ve). Fe (ec) Using Definition 5.4, we call a pair of puzzle pieces P , Pe for F , Fe corresponding if all Γ-itineraries of P coincide with the Γ-itinerariese of Pe. The definition then requires that the orbits of corresponding critical points c, ec travel through corresponding puzzle pieces at all depths. This also gives rise to the notion of corresponding critical fibers fib(c) and fib(ec). THE DYNAMICS OF COMPLEX BOX MAPPINGS 37

Observe that the degrees of the corresponding critical fibers for combinatorially equivalent box mappings must be equal (here, the degree of a fiber is the degree of F , respectively Fe restricted to a sufficiently deep puzzle piece around the fiber).

6. Statement of the QC-Rigidity Theorem and outline of the rest of the paper We are now ready to state the main result from [KvS], namely, the rigidity theorem for complex box mappings, stated there as Theorem 1.4, but with the explicit assumption that the complex box mappings are dynamically natural. Theorem 6.1 (Rigidity for complex box mappings). Let F : U → V be a dynamically natural complex box mapping that is non-renormalizable. Then (1) each point x ∈ K(F ) is contained in arbitrary small puzzle pieces. (2) If F carries a measurable invariant line field supported on a forward invariant set X ⊆ K(F ) of positive Lebesgue measure, then (a) there exists a puzzle piece J and a smooth foliation on J with the property that meas(X ∩ J ) = meas(J ) and the foliation is tangent to the invariant line field a.e. on J ; (b) the set Y = S F i(J ) intersects the critical set of F and each such critical i>0 point is strictly preperiodic. (3) Suppose Fe : Ue → Ve is another dynamically natural complex box mapping for which there exists a quasiconformal homeomorphism H : V → Ve so that (a) Fe is combinatorially equivalent to F w.r.t. H, and so in particular H(U) = Ue, (b) Fe ◦ H = H ◦ F on ∂U, i.e. F is a conjugacy on ∂U, Then F and Fe are quasiconformally conjugate, and this conjugation agrees with H on V\U. In the following sections, we sketch the proof of Theorem 6.1. Our aim is to make it more accessible to a wider audience.

Remark. Let us make several remarks regarding the statement of Theorem 6.1: (1) In [KvS], it was assumed that the box mapping in the statement of Theorem 6.1 is non-renormalizable and that all its periodic points are repelling. However, due to Lemma 5.2, it is enough to assume that the box mappings in question are non- renormalizable as the property that all periodic points are repelling then follows from that lemma. (2) In the language of fibers, shrinking of puzzle pieces stated in Theorem 6.1(1) is equivalent to triviality of fibers for points in the non-escaping set, i.e. fib(x) = {x} for every x ∈ K(F ). Note that if all the fibers are trivial, then c is combinatorially recurrent if and only if c ∈ ω(c). (3) There exist dynamically natural box mappings which have invariant line fields, see Proposition 13.1. These box mappings are rather special (and are the analogues of THE DYNAMICS OF COMPLEX BOX MAPPINGS 38

Latt`esrational maps in this setting), see Proposition 13.2. However, if a dynami- cally natural box mapping F is coming from a Yoccoz-type construction, namely, when the puzzles are constructed using rays and equipotentials (see Section 2.1), then each puzzle piece of F contains an open set which is not in K(F ) (these sets are in the Fatou set of the corresponding rational map). In this situation, conclu- sion (2a) of Theorem 6.1 cannot be satisfied, and thus such F carries no invariant line field provided F is non-renormalizable. (4) Conditions (3a) and (3b) should be understood in the following sense. Since, by assumption, H : V → Ve is a quasiconformal homeomorphism, it has a continu- ous homeomorphic extension Hˆ : cl V → cl Ve to the closures (see [AIM, Corollary 5.9.2]). This allows to use H in the definition of combinatorial equivalence (Defi- nition 5.4), and condition (3a) should be understood as ˆ Fe ◦ Hˆ = Hˆ ◦ Fˆ on ∂U, ˆ where Fˆ, Fe are the continuous extensions of F , Fe to the boundary. (5) If U only has a finite number of components, and each component of U, V, Ue, Ve is a quasi-disk and a combinatorial equivalence H : V → Ve as in Definition 5.4, then one can construct a new quasiconformal map H (by hand) which is a combinatorial equivalence and is a conjugacy on the boundary of U. Even if components of U, V, Ue, Ve are not quasi-disks, then one can slightly shrink the domains and still obtain a qc conjugacy between F, Fe on a neighborhood of their filled Julia sets. One also can often find a qc homeomorphism which is a conjugacy on the boundary using B¨ottcher coordinates when they exist. So in that case one essentially has that combinatorially equivalent box mappings F, Fe are necessarily qc conjugate. On the other hand, if U has infinitely many components, then it is easy to construct examples so that F,G are topologically conjugate while they are not qc conjugate. 6.1. Applications to box mappings which are not dynamically natural or which are renormalizable. In this subsection we discuss what can be said about rigidity of box mappings when we relax some of the assumptions in Theorem 6.1, that is if we do not assume that the box mapping is dynamically natural and/or non-renormalizable. First of all, the proof of Theorem 6.1(1) (given in Section 11) yields the following, slightly more general result: if F : U → V is a non-renormalizable complex box mapping satisfying the no permutation condition, then each point in Kwell-inside(F ) has trivial fiber. Furthermore, in [DS] it was shown that Theorem 6.1(1) can be extended to the following result [DS, Theorem C]: for a complex box mapping F : U → V and for every x ∈ K(F ) at least one of the following possibilities is true: (a) the fiber of x is trivial; (b) the orbit of x lands into a filled Julia set of some polynomial-like restriction of F ; (c) the orbit of x lands into a permutation component of V (i.e. there exists k > 1 and a component V of V which is also a component of U so that F k(V ) = V ); (d) the orbit of x converges to the boundary of V. THE DYNAMICS OF COMPLEX BOX MAPPINGS 39

It is easy to see that for dynamically natural complex box mappings the possibility (c) cannot occur, while points satisfying (d) have trivial fibers, and hence for dynamically natural mappings every point in its non-escaping set has either trivial fiber, or the orbit of this point lands in the filled Julia set of a polynomial-like restriction. These polynomial- like restrictions can be then further analyzed using the results on polynomial rigidity. For example, if the corresponding polynomials are at most finitely many times renormalizable and have no neutral periodic cycles, then the result in [KvS] tells that each point in the filled Julia set has trivial fiber, and hence the points in the whole non-escaping set of the original box mapping have trivial fibers. The discussion in the previous two paragraphs indicates the way how to extend Theo- rem 6.1(2), (3) for more general box mappings. For example, using Theorem 6.1(2), one can show that for every dynamically natural complex box mapping F , the support of any measurable invariant line field on K(F ) is contained in puzzle pieces with the properties described in (2a) and (2b) union the cycles of little filled Julia sets of possible polynomial-like renormalizations of F (compare [Dr]). Similarly, if for a pair of dynamically natural complex box mappings F and Fe that satisfy assumptions of Theorem 6.1(3) all polynomial-like renormalizations can be arranged in pairs F s : U → V , Fes : Ue → Ve so that U, Ue and V, Ve are the corresponding puzzle pieces and F s| , F s| are quasiconformally conjugate, then Theorem 6.1(3) can be used to U e Ue conclude that F and Fe are quasiconformally conjugate [DS]. Some conditions, for example in terms of rays for polynomial-like mappings, can be imposed in order to conclude rather than assume quasiconformal conjugation of the corresponding polynomial-like restrictions. The circle of ideas on how to “embed” results on polynomial dynamics into some ambient dynamics using complex box mappings, similar to the ones described in this subsection, is discussed in [DS] under the name rational rigidity principle. 6.2. Box mappings for which the domains are not Jordan disks. Let us now discuss Theorem 6.1 in the setting where we no longer assume that the components of U, V are Jordan disks, but instead assume that each component component U of U and V of V (i) is simply connected; (ii) when U b V then V \ U is a topological annulus; (iii) has a locally connected boundary. This means that U, V could for example be sets of the form D \ [0, 1). Such domains are often used for box mappings associated to real polynomial maps or to maps preserving the circle. In this setting - Parts (1) and (2) of Theorem 6.1 hold, provided Assumptions (i), (ii) are satisfied. The proof is identical as in the Jordan case, because these assumptions ensure that one still has annuli, V \ U, as before. - Part (3) of Theorem 6.1 holds, provided Assumptions (i), (ii), (iii) are satisfied and, moreover, the map H : V → Ve extends to a homeomorphism Hˆ : cl V → cl Ve. To see this, observe that the QC-Criterion, Theorem 8.6, also holds for simply con- nected domains Ω, Ωe provided that we now also assume that the homeomorphism THE DYNAMICS OF COMPLEX BOX MAPPINGS 40

ϕ:Ω → Ωe extends to a homeomorphismϕ ˆ: cl Ω → cl Ω.e (The proof of this version of the QC-criterion is the same as in [KSS1], since Assumption (iii) implies that we can apply the Carath´eodory Theorem to get that the Riemann mapping π : D → Ω extends to a continuous mapπ ˆ : cl D → cl Ω and thus the map ϕ:Ω → Ωe induces a homeomorphism ϕ: cl D → cl D. Because ϕ extends continuously to ∂Ω, the map ϕ has the additional property that if x, y ∈ ∂D andπ ˆ(x) =π ˆ(y), then ϕ(x) = ϕ(y). The K-qc map ψˆ: D → D so that ψˆ = ϕ on ∂D which is produced in the proof the QC-criterion, will now again have the same property that if x, y ∈ ∂D and πˆ(x) =π ˆ(y), then ψˆ(x) = ψˆ(y). Therefore, ψˆ produces a K-qc map ψ :Ω → Ωe which agrees with ϕ on ∂Ω.) Moreover, since H now is assumed to extend to a homeomorphism on the closure of the domains, the assumption that H is a con- jugacy on ∂U makes sense. Finally, the gluing theorems from Section 8.1 and the Spreading Principle go through without any further change.

6.3. Outline of the remainder of the paper. Let us briefly describe the structure of the upcoming sections. Theorem 6.1 contains three assertions about dynamically natural complex box mappings with all periodic points repelling: (1) on shrinking of puzzle pieces; (2) properties of box mappings with measurable invariant line fields; and (3) on quasiconformal rigidity (qc rigidity for short) of combinatorially equivalent, dy- namically natural box mappings, F : U → V and Fe : Ue → Ve. We outline the main ingredients that go into the proof of these assertions in Sections7 and8. For clarity and better referencing, these ingredients are shown in the diagram in Figure9, with arrows indicating dependencies between them. The main ingredients can be roughly divided into two groups: complex a priori bounds (discussed in Section7), and analytic tools (the Spreading Principle and the QC-Criterion, discussed in Section8). Complex bounds, see Theorem 7.1, gives us uniform control of the geometry of arbitrarily deep, combinatorially defined puzzle pieces. Complex bounds is the chief technical result needed to prove both shrinking of puzzle pieces (assertion (1)) and ergodic properties of box mappings and properties of mappings with measurable invariant line fields (assertion (2)). The shrinking of puzzle pieces is proved, assuming complex bounds, in Section 11. That the existence of an invariant line field implies the existence of a smooth foliation on some puzzle piece J is shown in Section 12 (again assuming complex bounds). In Section 13 we give an example of a dynamically natural box mapping with an invariant line field and show that such mappings have a very particular structure. We call these Latt`esbox mappings. We outline the proof of complex bounds in Section 10. It is important that the complex bounds are obtained for combinatorially defined puzzle pieces, since this ensures that corresponding puzzle pieces for combinatorially equivalent complex box mappings have good geometric properties simultaneously. This is a starting point of the proof of qc rigidity (assertion (3)) and it allows us to use the QC-Criterion (Theorem 8.6). THE DYNAMICS OF COMPLEX BOX MAPPINGS 41

In Section8, we continue the buildup to the explanation of the proof of qc rigidity. In that section, we explain the Spreading Principle and the QC-Criterion. The Spreading Principle, Theorem 8.5, allows one to obtain a partial quasiconformal conjugacy between a pair of combinatorially equivalent complex box mappings F and Fe outside of some puzzle neighborhood Y of Crit(F ), provided that one has a quasiconformal mapping from Y to the corresponding puzzle neighborhood Ye for Fe which respects the boundary marking, see Definition 8.3. Using both the QC-Criterion and the Spreading Principle, one can construct quasiconformal mappings that respect boundary marking between arbitrarily deep puzzle neighborhoods Y and Ye. In Section9, we combine all of the ingredients and explain the proof of qc rigidity. While we will assume that complex box mappings are dynamically natural, quite often we will only use part of the definition: to prove complex bounds around recurrent critical points, we will only need Definition 4.1(1); we will use Definition 4.1(2) to prove the Spreading Principle; and we will use Definition 4.1(3) to obtain complex bounds around non-recurrent critical points. In Section 14, we prove a Ma˜n´e-type result for dynamically natural complex box map- pings that was mentioned earlier in the introduction. THE DYNAMICS OF COMPLEX BOX MAPPINGS 42

Puzzle pieces shrink in diameter, QC rigidity of non-renormalizable existence of invariant line fields dynamically natural complex box mappings (Theorem 6.1(1)(2)) (Theorem 6.1(3))

Uniform control of dilatation Critical puzzle pieces shrink between arbitrary deep puzzle in diameter neighborhoods of critical sets (Theorem 7.1(1)) (Proposition 9.3)

Uniform control of geometry and extendibility for critical The Spreading Principle The QC-Criterion puzzle pieces (Theorem 8.5) (Theorem 8.6) (Theorem 7.1(2)(3))

Statements about a pair of maps

Uniform puzzle control for Uniform puzzle control, reluctantly recurrent and non-recurrent critical points persistently recurrent case (Lemmas 10.1 and 10.2) (Lemma 10.3) The Enhanced Nest for persistently recurrent critical points (Section 7.4) The Key Lemma (a priori bounds for persistently recurrent critical points, Lemma 10.4) The Kahn–Lyubich Covering Lemma (Lemma 7.10) Statements about a single map

Figure 9. The connection between the main ingredients that go into the proof of rigidity for non-renormalizable dynamically natural complex box mappings. The input ingredients are marked according to the following color scheme: the red boxes contain the ingredients coming from [KSS1], the green box contains the ingredient coming from [KvS], and finally, the blue box contains the ingredient proven in [KL1] (and used in [KvS]). THE DYNAMICS OF COMPLEX BOX MAPPINGS 43

7. Complex bounds, the Enhanced Nest and the Covering Lemma Complex bounds, also known as a priori bounds for a complex box mapping F : U → V concern uniform geometric bounds on arbitrarily small, combinatorially defined puzzle pieces. We will focus on two types of geometric control: moduli bounds and bounded geometry. This is the geometric control needed to apply the QC-Criterion in the proof of quasiconformal rigidity. It is not possible to control the geometry of every puzzle piece, and to obtain complex bounds, one must restrict ones attention to a chosen, combinatorially defined, nest of critical puzzle pieces, and show that one has complex bounds for the puzzle pieces in this nest. One example of a nest of puzzle pieces is the principal nest. In this section, we will discuss some of the geometric properties of the principal nest. We will also describe the Enhanced Nest, which was introduced in [KSS1]. It is combinatorially more difficult to define than the principal nest and it is only defined around persistently recurrent critical points, but it also provides much better geometric control on the puzzle pieces than the principal nest, and it is especially useful for mappings with more than one critical point.

7.1. Complex bounds for critical neighborhoods. Define Critac(F ) ⊂ Crit(F ) to be the set of critical points whose orbits accumulate on at least one critical fiber. Since it is a combinatorially defined set, it follows that for a combinatorially equivalent box mapping Fe the set Critac(Fe) consists of the corresponding critical points. Note that Crit(F )\Critac(F ) consists of what is usually called in real one-dimensional dynamics Misiurewicz critical points, i.e. critical points with non-recurrent fibers. The following theorem proven in [KSS1, Proposition 6.2] in the real case and extended to non-renormalizable complex maps in [KvS].

Theorem 7.1 (Shrinking neighborhoods with uniform control). If F : U → V is a non- renormalizable dynamically natural complex box mapping, then there exist a δ > 0 and an integer N such that the following holds. For every  > 0, there is a combinatorially defined nice puzzle neighborhood W of Crit(F ) with the following properties: (1) Every component of W has diameter < ; (2) Every component of W intersecting Critac(F ) has δ-bounded geometry; (3) The first landing map under F to W is (δ, N)-extendible with respect to Crit(F ). Moreover, if Fe is another non-renormalizable dynamically natural box mapping combina- torially equivalent to F , then the same claims as above hold for Fe and the corresponding objects with tilde.

Remark. In fact, for persistently recurrent critical points, we have better control. There 0 exists a beau constant δ > 0 such that if a component Wc of W is a neighborhood of a 0 persistently recurrent critical point, then Wc has δ -bounded geometry and the first landing 0 map to Wc is (δ ,N)-extendible with respect to Crit(F ). The constant N is bounded by some universal function depending only on the number and degrees of the critical points of F . THE DYNAMICS OF COMPLEX BOX MAPPINGS 44

Let us recall a notion of (δ, N)-extendibility defined in [KSS1, page 775]. Let δ > 0 and 0 N ∈ N. Let V ⊃ V be a pair of nice open puzzle neighborhoods of some set A ⊂ Crit(F ), 0 and B be any subset of Crit(F ). For a ∈ A, denote Va = Compa V ; similarly for Va. We 0 say that the first landing map LV to V is N-extendible to V w.r.t. B if the following hold s 0 −s 0 (see Figure 10): if F : U → Va is a branch of LV , and if U = CompU (F (Va)), then  j 0 j # 0 6 j 6 s − 1: (F (U ) \ F (U)) ∩ B 6= ∅ 6 N. 0 Moreover, LV is (δ, N)-extendible w.r.t. B if there exists a puzzle piece Va ⊃ Va for every 0 S 0 a ∈ A such that mod(Va \ V a) > δ and LV is N-extendible to a∈A Va w.r.t. B. Remark. Note that this definition is slightly more demanding than the one from [KSS1] 0 because here we require that the set Va to be a puzzle piece rather than a topological disk. The reason this can be done is because we are able to prove this sharper statement in Lemma 10.3 as now we can rely on the estimate provided in Proposition 10.5 from [KvS]. In [KSS1] such an estimate is not available.

Figure 10. A schematic drawing of relevant annuli in the definition of (δ, N)-extendibility: δ controls the moduli of green annuli, N controls the s number of critical points in the blue annuli for each branch F : U → Va of the first landing map LV . The set V is the union of solid orange disks; the set A is indicated with black dots, the set B is indicated with blue crosses. THE DYNAMICS OF COMPLEX BOX MAPPINGS 45

We will sketch the proof of Theorem 7.1 in Section 10. There are two main ingredients to this proof. The first one is the construction from [KSS1] of the around persistently recur- rent critical points. The second one is the Covering Lemma of Kahn and Lyubich [KL1], see [KvS]. In [KSS1] other methods were used in the real case because [KL1] was not avail- able, and also to deal with the infinitely renormalizable case. Before going on to discuss these ingredients in greater detail, we will describe the geometry of the principal nest.

7.2. The principal nest. Let c ∈ Crit(F ) be a critical point. The principal nest (around c) is a nested collection of puzzle pieces V 0 ⊃ V 1 ⊃ V 2 ⊃ ... such that

0 i+1 i c ∈ V , and V = Lc V for every integer i > 0. Note that we can start the principal nest from an arbitrary strictly nice puzzle piece V 0 containing c. Clearly, if c is combinatorially non-recurrent, then the principal nest is finite. Otherwise, the nest consists of infinitely many puzzle pieces. Moreover, each V n contains fib(c), and hence all the critical points in fib(c). Associated to such the principal nest we have a sequence of return times (pn) defined n+1 −pn n n as V = Compc F (V ); pn is the first time the orbit of c returns back to V . Since the mapping from V n+1 to V n is a first return mapping, and puzzle pieces are nice sets, j n+1 j0 n+1 0 the sets F (V ),F (V ) are pairwise disjoint for 0 < j < j 6 pn. Thus all the maps F pn : V n+1 → V n have uniformly bounded degrees.

7.2.1. Geometry of the principal nest. Let us describe some of the results concerning the geometry of the principal nest for unicritical mappings. Suppose that F : U → V is a dynamically natural complex box mapping with the properties that •V consists of exactly one component, • there exists exactly one component U0 of U that is mapped as a degree d branched covering onto V which is ramified at exactly one point, and • every component U of U, U 6= U0 is mapped diffeomorphically onto V. Let V 0 = V and V 0 ⊃ V 1 ⊃ V 2 ⊃ ... be the principal nest about the critical point c of F . There are two combinatorially distinct types of returns F pn : V n+1 → V n: central and non-central. We say that the return mapping from V n+1 to V n is central if F pn (c) ∈ V n+1. Otherwise the return is called non-central. For quadratic mappings, the following result is fundamental.

Theorem 7.2. [GS, Lyu2, Decay of Geometry] Let nk denote the subsequence of levels in p n n −1 the principal nest so that F nk−1 : V k → V k is a non-central return. There exists a constant C > 0 so that  n +1 mod V nk \ V k > Ck.

 THE DYNAMICS OF COMPLEX BOX MAPPINGS 46

For example, by the Koebe Distortion Theorem this result implies that the landing maps to V nk+1 become almost linear as k → ∞. This result played a crucial role in the proofs of quasiconformal rigidity and density of hyperbolicity in the quadratic family. For higher degree maps, even cubic mappings with two non-degenerate critical points and unicritical cubic mappings, Decay of Geometry does not hold; however, the geometry is bounded from below. The following result required new analytic tools — the Quasiad- ditivity Law and Covering Lemma — of [KL1].

Theorem 7.3 ([AKLS, ALS]). Let nk denote the subsequence of levels in the principal p n n −1 nest so that F nk−1 : V k → V k is a non-central return. There exists a constant δ > 0 so that  n +1 mod V nk \ V k > δ.  Even for unicritical mappings, this difference makes the study of higher degree mappings quite a bit different from the study of unicritical mappings with a critical point of degree 2. n+1 In general, there is no control of the moduli of the annuli V n \ V when the returns are central. Observe that when the return mapping from V n+1 to V n is central, we have that the return time of c to V n+1 agrees with the return time of c to V n, and hence the annulus V n+1 \ V n+2 is mapped by a degree d covering map onto the annulus V n \ V n+1. Hence  n+2 1  n+1 mod V n+1 \ V = mod V n \ V . d Consequently, when there is a long cascade of central returns; that is, when F pn (c) ∈ V n+k n+k n+1+k 1 n n+1 for k large, we have that mod(V \ V ) = dk mod(V \ V ) may be arbitrarily small. In general, it is impossible to bound the length of long central cascades, so these moduli can degenerate. ∞ For unicritical mappings, the complex a priori bounds imply that if {pn}n=0 is the sequence of levels in the principal nest so that F pn−1 : V n → V n−1 is non-central, then the mapping F pn : V pn+1 → V pn is (δ(n), 1)-extendible, where δ(n) → ∞ at least linearly as n → ∞ when the critical point has degree 2, and δ(n) is bounded from below for higher degree mappings. Example: Puzzle pieces in the principal nest do not have bounded geometry in general. De- spite having moduli bounds, one cannot expect to have bounded geometry for the principal nest, even at levels following non-central returns. For example, suppose that F : U → V is real-symmetric box mapping, and that F pn : V n+1 → V n has a long central cascade, so p n+k p n+1 10 that F n (0) ∈ V , and that F n (V ) ∩ R 3 0. Then for any  > 0, as the length of the cascade tends to infinity the puzzle piece V n+k is contained in an  · diam(V n+k)- neighborhood of its real trace. The reason for this is that as k → ∞, the puzzle pieces in the principal nest converge to the Julia set of the return mapping, which in this case,

10Such a return is called high. THE DYNAMICS OF COMPLEX BOX MAPPINGS 47 is close to the real trace of V n+k as the return mapping is, up to rescaling, is close to z 7→ z2 − 2. Hence in general one cannot recover δ-bounded geometry for subsequent puzzle pieces, even if the return mappings are non-central.

Example: the Fibonacci map. Let us assume that F : U → V is a real-symmetric, unicritical complex box mapping with quadratic critical point at 0 and Fibonacci combinatorics (see Figure 11, left); that is, for each V n we have that there are exactly two components V n+1 n n and V1 of the domain of the first return map to V that intersect the postcritical set of F . pn n Moreover, we assume that for all n, F (0) ∈ V1 , so that all of the first return mappings n n to the principal nest are non-central, that the return mapping from V1 to V is given by p p n+1 F n−1 | n , and that restricted to the real trace F n (V ∩ ) 3 0; that is, these return V1 R mappings are all high.

Figure 11. (Left) The Fibonacci combinatorics. (Right) The actual shape of the n-th puzzle piece V n in the principal nest of the Fibonacci mapping of degree two.

For (quadratic) Fibonacci mappings, Decay of Geometry together with the precise de- scription of the combinatorics implies that the puzzle pieces V n converge to the shape of the filled Julia set of z 7→ z2 − 1, the basilica [Lyu6]. In particular, these puzzle pieces become badly pinched and the hyperbolic distance between 0 and F pn (0) in V n tends to infinity, see Figure 11, right. Consequently, these puzzle pieces are not δ-free (see Sec- tion 1.5 to recall the definition). However, the return mappings are (K, 1)-extendible for K-arbitrarily large at sufficiently deep levels, and there exists δ > 0, independent of n, so that the puzzle pieces have δ-bounded geometry.

7.3. Inducing persistently recurrent box mappings. Recall that a box mapping F 0 : U 0 → V0 is induced from F : U → V if U 0 and V0 consist of puzzle pieces of F and the branches of F 0 are the iterates of some branches of F . THE DYNAMICS OF COMPLEX BOX MAPPINGS 48

Definition 7.4 (Persistently recurrent box mapping). We say that a non-renormalizable dynamically natural box mapping F is persistently recurrent if every critical point of F is persistently recurrent and its orbit accumulates at every critical fiber of F . The latter is equivalent to the property that Crit(F ) = [c] for some, and hence for every c ∈ Crit(F ) (see notation in Section 5.2). The next lemma shows that starting from an arbitrary box mapping F (non-renormalizable and dynamically natural) and a persistently recurrent critical point c ∈ Crit(F ) we can extract an induced box mapping F 0 that is persistently recurrent and contains c, and hence fib(c), in its non-escaping set. More precisely: Lemma 7.5 (Inducing persistently recurrent box mappings). Let F be a dynamically natural box mapping that is non-renormalizable, and c ∈ Crit(F ) be a persistently recurrent critical point of F . Then there exists an induced persistently recurrent box mapping F 0 such that c ∈ Crit(F 0) ⊂ Crit(F ). Proof. By definition of persistent recurrence, the set [c] consists only of persistently recur- rent critical points. Choose an integer m large enough so that each critical puzzle piece of 0 depth at least m contains a single critical fiber. Let Vn be the union of all puzzle pieces of F of depth n > m intersecting [c]; this is a finite union of critical puzzle pieces containing 0 0 only persistently recurrent critical fibers. Let Un be the union of the components of R(Vn) 0 0 0 0 intersecting [c], and define a box mapping Fn : Un → Vn to be the restriction to Un of the 0 first return map to Vn under F . 0 Let us show that there exists an n(> m) such that the only critical points of Fn are the ones in [c]. This will be our desired box mapping F 0. 0 0 Assume the contrary, and let Un ⊂ Un and Vn ⊂ Vn be a pair of components containing, 0 00 0 respectively, c ∈ [c] and c ∈ [c], and such that Crit(F ) ∩ Un 6⊂ [c]. This means that if s F : Un → Vn, s = s(n) is the corresponding branch of the first return map, then one of 2 s−1 000 the puzzle pieces in the sequence F (Un),F (Un),...,F (Un) contains a c ∈ Crit(F ) with is either non-persistently recurrent critical point or persistently recurrent critical point with [c] ∩ [c000] = ∅. If this is true for every n, then the orbits of c0, c00 and c000 must all visit arbitrarily deep puzzle pieces around each other. In particular, c000 must be combinatorially recurrent and [c000] must intersect [c0]. On the other hand, c000 cannot be reluctantly recurrent: the critical point c0 will then have only finitely many children, contrary to the definition of reluctant recurrence. We arrived to a contradiction.  7.4. The Enhanced Nest. In this subsection, we review the construction and properties of the Enhanced Nest of puzzle pieces defined around a chosen persistently recurrent critical point. Throughout Section 7.4, unless otherwise stated we will assume that all complex box mappings are dynamically natural, non-renormalizable, and persistently recurrent. By Lemma 7.5 we know that we can always induce such a mapping starting from a given (dynamically natural non-renormalizable) box mapping. The construction of the Enhanced Nest is based on the following lemma [KSS1, Lemma 8.2] (see Figure 12, left for an illustration): THE DYNAMICS OF COMPLEX BOX MAPPINGS 49

Lemma 7.6. Let F be a persistently recurrent box mapping, c ∈ Crit(F ) be a (persis- tently recurrent) critical point, and I 3 c be a puzzle piece. Then there exists a com- binatorially defined positive integer ν with F ν(c) ∈ I such that the following holds. If −ν j U0 := Compc (F (I)), and Uj := F (U0) for 0 < j 6 ν, then 2 (1) # {0 6 j 6 ν − 1: Uj ∩ Crit(F ) 6= ∅} 6 b , where b is the number of critical fibers of F (which all intersect orb(c)); −ν  (2) U0 ∩ PC(F ) ⊂ Compc F LF ν (c)(I) .  This lemma allows us to define the following pullback operators A and B. Let c ∈ Crit(F ) be a persistently recurrent critical point, and for a critical puzzle piece I 3 c, let ν = ν(I) be the smallest possible integer with the properties specified by Lemma 7.6. Define −ν  A(I) := Compc F LF ν (c)(I) , −ν  B(I) := Compc F (I) . Let us comment on the meaning of these operators. First of all, these operators depend on persistently recurrent critical point (or to be more precise, on the corresponding fiber, see Section 5.1). Once the critical point is fixed, say c, the operators A: I 7→ A(I) and B : I 7→ B(I) produce for any given puzzle piece I 3 c its pullbacks A(I) 3 c, respectively B(I) 3 c so that: • the degrees of the corresponding maps F s : A(I) → I and F ν : B(I) → I are bounded only in terms of the local degrees of the critical points of F , and hence are independent of I (we can thus speak of the bounds on degrees assigned to the operators); this property follows by Lemma 7.6(1) (note that U0 = B(I) in that lemma); • the puzzle piece A(I) has some “external space” free of the critical and postcritical set of F (denoted earlier as PC(F )), while the puzzle piece B(I) has some “internal space” disjoint from PC(F )11, see Figure 12, left. This property is Lemma 7.6(2). It is immediate from the last property that if we apply B after A, then for every c ∈ I the puzzle piece BA(I) will be the pullback of I of uniformly bounded degree and will have an annular “buffer” around its boundary free of PC(F ), see Figure 12 right. This is a key step in the construction of the Enhanced Nest below. In fact, these buffers can be written explicitly in terms of A and B. If we set P = BA(I),P + = BB(I) and P − = AA(I), then the annuli P + \ cl (P ) and P \ cl (P −) are disjoint from PC(F ). In Lemma 10.4, we will show that there is a uniform bound on the moduli of such “buffers” for the elements in the Enhanced Nest. For technical reasons that are apparent in the proof of Lemma 7.8, it is not enough to produce a nest only by applying A and B starting with some puzzle piece I. In order to

11One might interpret the notation as follows: the operator A gives puzzle pieces with some space free of PC(F ) Above the boundary of the piece, while the operator B gives pieces with some space free of PC(F ) Below the boundary of the piece. THE DYNAMICS OF COMPLEX BOX MAPPINGS 50

LF ν (c)(I) I F ν(c) I A F ν

B

c c | {z } ⊂PC(F ) F ν

A(I) A(I)

B(I) = U0 BA(I)

Figure 12. The role of the pullback operators A and B: the operator A creates some “external” space free of PC(F ), while B creates some “inter- nal” space free of PC(F ). The annuli disjoint from PC(F ) are shaded in yellow. The figure on the left illustrates the construction in Lemma 7.6; here, elements of the set PC(F ) are marked with crosses and dots. The figure on the right illustrates the result of successive applications of A and B: the puzzle piece BA(I) has an annular “buffer” free of PC(F ) around its boundary. construct an effective nest that goes fast enough in depth, we have to use several times the following smallest successor construction. Given a puzzle piece P 3 c, by a successor of P we mean a puzzle piece containing c ˆ ˆ 0 of the form Compc L(Q), where Q is a child of Compc0 L(P ) for some c ∈ Crit(F ). If c is persistently recurrent, then by definition each critical puzzle piece I 3 c has the smallest (i.e. deepest) successor, which we denote by Γ(I). By construction, the corresponding return map F s : Γ(I) → I has degree bounded only in terms of the local degrees of the critical points of F , and hence independent of I. Definition 7.7 (The Enhanced Nest). The Enhanced Nest (around a persistently recurrent critical point c, or, equivalently, around a persistently recurrent critical fiber fib(c) of a persistently recurrent box mapping F ) is a nested sequence I0 ⊃ I1 ⊃ I2 ... of puzzle pieces such that

5b c ∈ I0, and Ii+1 = Γ BA(Ii) for every integer i > 0, where b is the number of critical fibers intersecting orb(c). THE DYNAMICS OF COMPLEX BOX MAPPINGS 51

By the discussion above, In+1 is a pullback of In and the return map pn F : In+1 → In has degree bounded independently of n. Moreover, each element In in the Enhanced Nest + − − + comes with a pair of puzzle pieces In ,In nested as In ⊂ In ⊂ In so that the annuli + − In \ cl (In) and In \ cl (In ) are disjoint from PC(F ). This particular construction of the Enhanced Nest is chosen because of the following lemma ([KSS1, Lemma 8.3]): Lemma 7.8 (Transition and return times for the Enhanced Nest). Let c ∈ Crit(F ) be a persistently recurrent critical point, and (In)n>0 is the Enhanced Nest around c. If pn is the transition time from In+1 to In, and r(In+1) is the return time of In+1 to itself, i.e. r(In+1) is the minimal positive integer r such that there exists a point x ∈ In+1 with r F (x) ∈ In+1, then (1) 3r(In+1) > pn; (2) pn+1 > 2pn.  This lemma is one of the main ingredients in the proof of the Key Lemma (Lemma 10.4). Let us comment on the meaning of conditions (1) and (2) in Lemma 7.8. s Condition (1) implies the following: if F : A → In+1 is an arbitrary branch of the first return map to In+1 under F , then s > r(In+1) > pn/3. As for condition (2), first observe that (pn)n>0 is a monotonically increasing sequence of integers. Hence, by (2), pn+1 > 2pn > pn + pn−1. This means that this sequence grows faster than the sequence of Fibonacci numbers (with the same starting pair of values). For the classical Fibonacci numbers fn, that is when fn+1 = fn + fn−1, f0 = f1 = 1, the following property is well-known: n n X X fn+2 = f1 + fi = 1 + fi. i=0 i=0

Along the same lines one can show that for our sequence (pi)i>0 for every n > 0 we have n X pn+2 > p1 + pi, i=0 or more generally, for every n > m > 0, n n X X (7.1) pn+2 > pm+1 + pi > pi. i=m i=m Pn t t If we write t := i=m pi, then F (In+1) = Im and the degree of the map F : In+1 → Im can be arbitrarily big. However, restricted to In+3 this map has degree bounded indepen- dently of n and m. This follows from (7.1): it takes additional pn+2 − t > 0 iterates of F to t t map F (In+3) over In+2, and hence the degree of F |In+3 is bounded above by the degree p of F n+2 : In+3 → In+2, which is uniformly bounded (see Figure 13). This mechanism is in the core in the proof of Proposition 10.5, as it allows to apply the Covering Lemma, which we discuss next. THE DYNAMICS OF COMPLEX BOX MAPPINGS 52

Figure 13. The inequality for the transition times in the Enhanced Nest t Pn implies that for every n > m > 0, the puzzle piece F (In+3), t = i=m pi is contained in some entry domain Lx(In+2) to In+2 under F , and the degree t of the restriction F |In+3 is bounded independently of n and m, while the t degree of F : In+1 → Im does not.

7.5. The Covering Lemma. The second main ingredient that go into the proof of the Key Lemma (Lemma 10.4) is the Kahn–Lyubich Covering Lemma, which we discuss is this subsection. 0 0 0 Let A ⊂ A ⊂ U, B ⊂ B ⊂ V be open topological disks in C. Denote by f :(A, A ,U) → (B,B0,V ) a holomorphic branched covering map f : U → V that maps A0 onto B0 and A onto B. One of the standard results on distortion of annuli under branched coverings is the following lemma; its proof is based on the Koebe Distortion Theorem. Lemma 7.9 (Standard distortion of annuli). Let A ⊂ U and B ⊂ V be open topological disks, and f :(A, U) → (B,V ) be a holomorphic branched covering map. Suppose that the degree of f is bounded by D > 0. Then −1 mod(U \ A) > D · mod(V \ B).  In this lemma, once the degree D starts to grow, we lose control on the modulus of the pullback annulus. However, in the situation when the annulus U \ A is nearly degenerate some portion of the modulus under pullback can still be recovered, provided that the restriction of f to some smaller disk inside U is of much smaller degree compared to D. This situation is often the case in complex dynamics, for example, when one considers a pair P ⊂ Q of critical puzzle pieces such that F s(P ) = Q for some large s: the degree of s F |P can be arbitrary high, but restricted to some small neighborhood of the critical point in P the degree is much smaller. This was made precise by Kahn and Lyubich [KL1] in the following lemma: THE DYNAMICS OF COMPLEX BOX MAPPINGS 53

Lemma 7.10 (Covering Lemma). For any η > 0 and D > 0 there is  = (η, D) > 0 such that the following holds: Let A ⊂ A0 ⊂ U and B ⊂ B0 ⊂ V be a triple of nested and open topological disks, and let f :(A, A0,U) → (B,B0,V ) be a holomorphic branched covering map. Suppose the degree of f is bounded by D, and the degree of f|A0 is bounded by d. Then

−1 0 −2 mod(U \ A) > min(, η · mod(B \ B), Cηd · mod(V \ B)), where C > 0 is some universal constant.  Let us make some comments on the statement of the Covering Lemma. The parameter  in this lemma controls the degeneration of the annulus U \ A: either U \ A has modulus at least , or mod(U \ A) <  and we are in a nearly degenerate regime. Note that the extend to which we have to degenerate to “enter” this regime depends on η and the large degree D. Once we are in the nearly degenerate regime, then the Covering Lemma says that under the pullback we either recover at least η−1-portion of the modulus of B0 \ B (called the collar in the terminology of [KL1]), or mod(U \ A) is even smaller than η−1 · mod(B0 \ B) and in this even more degenerate regime we recover at least Cηd−2-portion of mod(V \ B).

U V D : 1

0 A 0 d : 1 B

A B

Figure 14. The annuli in the Covering Lemma.

In the real-symmetric case, the Covering Lemma was proven in a sharper form in [KvS, Lemma 9.1].

8. The Spreading Principle and the QC-Criterion In this section we discuss two quite general tools that are used to prove qc rigidity: the Spreading Principle and the QC-Criterion. We start with some auxiliary results on quasiconformal mappings and their gluing properties. THE DYNAMICS OF COMPLEX BOX MAPPINGS 54

8.1. Quasiconformal gluing. The following lemma is well-known (see for example [DH1, Lemma 2]), and is due to Rickman [Ri] (sometimes it is also called Bers’ Sewing Lemma and attributed to Bers).

Lemma 8.1 (Sewing Lemma). Let U ⊂ C be an open topological disk and E ⊂ U be a relatively compact set. Let ϕi : U → ϕi(U), i ∈ {1, 2} be a pair of homeomorphisms such that

(1) ϕ1 is K1-quasiconformal, (2) ϕ2|U\E is K2-quasiconformal, (3) ϕ1|∂E = ϕ2|∂E. Then the map Φ: U → Φ(U) defined as ( ϕ1(z) if z ∈ E Φ(z) = ϕ2(z) if z ∈ U \ E is a K-quasiconformal homeomorphism with K = max(K1,K2). Moreover, ∂Φ = ∂ϕ1 almost everywhere on E.  We will need the following corollary to the Sewing Lemma, which is immediate if in the statement below the number of disks is finite; if this number is not finite, then the argument is slightly more delicate (see also the remark below).

Lemma 8.2 (Countable gluing lemma). Let V ⊂ C be an open topological disk and ψ : V → ψ(V ) be a K-quasiconformal homeomorphism. Let Bi ⊂ V , i ∈ I be an at most countable collection of closed Jordan disks with pairwise disjoint interiors. Suppose 0 there exist homeomorphisms ϕi : Bi → ϕi(Bi) that are uniformly K -quasiconformal in the interior of Bi and match on the boundary with ψ, i.e. for each i ∈ I,

ψ|∂Bi = ϕi|∂Bi . Then the map Φ: V → Φ(V ) defined as ( ϕi(z) if z ∈ Bi Φ(z) = ψ(z) if z ∈ V \ ∪i∈I Bi is a max(K,K0)-quasiconformal homeomorphism.

Remark. In general, if I is infinite and one assumes that ψ and all ϕi are just homeo- morphisms, then the map Φ need not be even a homeomorphism. Lemma 8.2 allows us in the statement of qc rigidity for box mappings to avoid the assumption that the domains of U shrink to points, which was imposed in [Lyu2] and [ALdM]. Under that assumption in those papers [Lyu2, Lemma 11.1] was used instead of the Gluing Lemma above.

Proof of Lemma 8.2. Exhaust I by finite sets In. For each n, define ( ϕi(z) if z ∈ Bi for some i ∈ In Φn(z) := ψ(z) if z ∈ V \ ∪i∈In Bi THE DYNAMICS OF COMPLEX BOX MAPPINGS 55

Since In is finite and each ϕi, i ∈ In matches with ψ on the boundary of Bi, the map Φn is a homeomorphism of V onto its image. By the Sewing Lemma (Lemma 8.1) applied to E = ∪i∈In ∂Bi, ϕ1 = ψ, and ϕ2 = Φn, it follows that ( ψ(z) if z ∈ ∪i∈I ∂Bi z 7→ n Φn(z) if z ∈ V \ ∪i∈In ∂Bi

0 is a max(K,K )-quasiconformal homeomorphism. The latter map is clearly equal to Φn. 0 In this way, we obtained a sequence (Φn)n>0 of max(K,K )-quasiconformal maps on V . By compactness of quasiconformal maps of uniformly bounded dilatation, this sequence has a sub-sequential limit. Moreover, since the maps in the sequence agree on more and more disks Bi, all sub-sequential limits are the same and equal to the map Φ. The conclusion of the lemma follows. 

8.2. Lifts of combinatorial equivalence and the boundary marking. Let F : U → V and Fe : Ue → Ve be a pair of complex box mappings that are combinatorially equivalent in the sense of Definition 5.4 w.r.t. to some K-quasiconformal homeomorphism H : V → Ve such that H(U) = Ue and Fe ◦ H = H ◦ F on ∂U for each component U of U. By Remark4 after Theorem 6.1, the latter condition means that if Hˆ : cl V → cl Ve is a continuous homeomorphic extension of H to the closures, then

ˆ ˆ ˆ ˆ Fe ◦ H|∂U = H ◦ F |∂U ˆ for each component U of U; here Fˆ and Fe are the corresponding homeomorphic extensions of F and Fe to the closures. Let P and Pe be a pair of corresponding puzzle pieces for F and Fe in the sense of Section 5.4. Definition 8.3 (Boundary marking). We say that a homeomorphism h: P → Pe respects the boundary marking (induced by H) if h extends to a homeomorphism hˆ between cl P ˆ ˆ and cl Pe and h|∂P agrees with the corresponding lift of H; that is, if k > 0 is the depth of P , then ˆ k ˆ ˆ ˆk Fe ◦ h|∂P = H ◦ F |∂P . Let K be the quasiconformal dilatation of H. The definition of combinatorial equivalence then gives us a K-qc homeomorphism that respects the boundary marking between any two corresponding puzzle pieces of depths 0 and 1. The next lemma is a simple observation that shows that such H gives rise to a K0-qc homeomorphism that respects the boundary 0 marking for a pair of corresponding arbitrary deep puzzle pieces; of course, K > K can be arbitrary large.

Lemma 8.4 (Starting qc maps respecting boundary). Let P , Pe be a pair of corresponding 0 puzzle pieces of depth k > 0 for F , Fe. Then there exists a K -quasiconformal homeomor- phism h: P → Pe that respects the boundary marking induced by H. THE DYNAMICS OF COMPLEX BOX MAPPINGS 56

Proof. Let V be the component of V so that F k(P ) = V . If F k : P → V is a conformal isomorphism, then, by combinatorial equivalence (Definition 5.4), Fek : Pe → Ve is also a conformal isomorphism. In this situation, we can define h := Fe−k ◦ H ◦ F k for the branch Fe−k that maps Ve over Pe. Clearly, h is a K-qc map that respects the boundary marking. k Suppose v1, . . . , v` ∈ V are the critical values of the map F : P → V . By combinatorial equivalence, the map Fek : Pe → Ve must have the same number of critical values with the same branching properties (i.e. if c ∈ P is a critical point of local degree s > 2 and so that k F (c) = vi, then the corresponding critical point ec must have local degree s and is mapped k by Fe to the corresponding critical value vei). −1 Choose an arbitrary smooth diffeomorphism L: V → V so that L(vi) = H (vei) and so that L = id in some small neighborhood of ∂V . Then H ◦ L: V → Ve is a K0-qc homeomorphism that maps the critical values of F k| to the critical values of F k| and P e Pe respects the boundary marking. Such homeomorphism can be now lifted as above: define h = Fe−k ◦ H ◦ L ◦ F k (for appropriate choices of inverse branches). By construction, h is K0-qc homeomorphism between P and Pe that respects the boundary marking, as required.  8.3. The Spreading Principle. The Spreading Principle is a dynamical tool which makes it possible to construct quasiconformal conjugacies outside of some puzzle neighborhood Y of Crit(F ) between two mappings F : U → V and Fe : Ue → Ve provided that one has an initial quasiconformal mapping H : V → Ve which is a combinatorial equivalence and a conjugacy on ∂U between F and Fe, and a quasiconformal mapping on each component P of Y to the corresponding component Pe of Ye that respects the boundary marking (in the sense of Definition 8.3). In the context of polynomials, the Spreading Principle was proven in [KSS1, Section 5.3, page 769] and [KvS]. Below we state and prove the corresponding version for dynamically natural box mappings. Our statement is slightly more general than the one in [KSS1, KvS] as it allows to take into account some pre-existing conjugacy on some part of the postcritical set. Theorem 8.5 (Spreading Principle). Let F : U → V and Fe : Ue → Ve be a pair of dynami- cally natural complex box mappings and H : V → Ve be a K-quasiconformal homeomorphism that • provides a combinatorial equivalence between F and Fe, is a conjugacy on the bound- ary of U, and moreover, it i • conjugates F to Fe on XS := {F (fib(c)): c ∈ S, i > 0} for some subset S ⊂ Crit(F ) (possibly empty).

Let Y be a nice puzzle neighborhood of Crit(F ) \ S such that XS ∩ Y = ∅, and let Ye be the corresponding neighborhood of Crit(Fe) \ Se for Fe. Further suppose that there exists a K0-quasiconformal homeomorphism ϕ: Y → Ye that respects the boundary marking induced by H. Then ϕ extends to a max(K,K0)-quasiconformal homeomorphism Φ: V → Ve such that: THE DYNAMICS OF COMPLEX BOX MAPPINGS 57

(1) Φ = ϕ on Y; (2) for each z ∈ U \ Y, Fe ◦ Φ(z) = Φ ◦ F (z);

(3) Φ|V\Lˆ(Y) is K-quasiconformal; (4) Φ(P ) = Pe for every puzzle piece P that is not contained in Lˆ(Y), and Φ: P → Pe respects the boundary marking induced by H.

Proof. Denote by Pn the collection of puzzle pieces of depth n > 0 for F . Since XS ∩ Y = ∅ and XS is forward invariant under F , it follows that XS ∩ Lˆ(Y) = ∅; the same is true for the corresponding sets X and Y. Moreover, there exists a puzzle eSe e neighborhood X of S consisting of puzzle pieces of the same depth, say m, with the property that m is larger than the depths of the components of Y and X ∩ Y = ∅. Note that Y ∪ X is a nice set by construction. Let C be a collection of all critical puzzle pieces that are not contained in Y but do intersect Crit(F ) \ S. This is a finite set. Using Lemma 8.4, for each C ∈ C let us pick a quasiconformal map hC : C → Ce that respects the boundary marking induced by H; suppose K0 is the maximum over the quasiconformal dilatations of hC for C ∈ C. We now want to construct a quasiconformal homeomorphism ψ : X → Xe that respects k the boundary marking. Let Q be a component of X , and k > 0 be minimal so that F (Q) either is a component of V, or is in C. Now define a map ψQ : Q → Qe by the formula k k Fe ◦ ψQ = f ◦ F , where (H if F k(Q) is a component of V, (8.1) f = k hF k(Q) if F (Q) ∈ C. Since none of the puzzle pieces in the sequence Q, F (Q),...,F k(Q) intersects Y (as XS ∩ Y = ∅), and F is conjugate to Fe on XS, we conclude that the map ψQ is a well- defined quasiconformal homeomorphism with dilatation max(K,K0). By construction, ψQ respects the boundary marking induced by H. We define ψ : X → Xe component-wise using the maps ψQ. Note that the dilatation of ψ does not depend on the choice of the puzzle neighborhood X , and as the depth of the neighborhood X tends to ∞, the set where the dilatation of ψ is equal to K0 shrinks to zero (in measure). Finally, let us adjust ϕ, Y and Y0 as follows. Set Y0 := Y ∪ X and Ye0 := Ye ∪ Xe, and define (ϕ(z) if z ∈ Y, ϕ0(z) = ψ(z) if z ∈ X .

Defined this way, ϕ0 : Y0 → Ye0 is a K00-quasiconformal homeomorphism between nice puzzle 00 0 neighborhoods of Crit(F ) and Crit(Fe), with K := max(K,K ,K0), and this homeomor- phism respects the boundary marking. ˆ 0 Let Q be a component of L(Y ), and k > 0 be the landing time of the orbit of Q to Y0. Since Crit(F ) ⊂ Y0, the map F k : Q → F k(Q) is a conformal isomorphism. Therefore, THE DYNAMICS OF COMPLEX BOX MAPPINGS 58

0 00 0 we can pull back ϕ and define a K -quasiconformal homeomorphism ϕQ : Q → Qe by the k 0 0 k 0 formula Fe ◦ ϕQ = ϕ ◦ F . Since ϕ respects the boundary marking induced by H, so does 0 ϕQ. Let us now inductively define a nested sequence of sets Y0 ⊃ Y1 ⊃ Y2 ... so that:

• Y0 to be the union of all puzzle pieces in P0; • Yn+1 is the subset of Yn consisting of puzzle pieces of depth n + 1 that are not components of Lˆ(Y0).

For each puzzle piece Q ∈ Pn that is contained in Yn, it follows that none of the pieces 0 in the orbit Q, F (Q),... can lie in Y (because otherwise Q would lie in Y`−1 \ Y` for k some ` 6 n). Therefore, the exists the minimal k = k(Q) > 0 so that F (Q) either is in C, or is a component of V. By a similar reasoning as above, in both cases the map F k : Q → F k(Q) allows to pull back homeomorphisms consistently and to define k k a homeomorphism HQ : Q → Qe by the formula Fe ◦ HQ = f ◦ F , where f is as in (8.1). Defined this way, the map HQ is max(K,K0)-quasiconformal map that respects the boundary marking induced by H. X X We proceed by inductively defining a sequence of homeomorphisms (Φn )n>0,Φn : V → Ve X as follows. Set Φ0 = H, and for each n > 0 define ΦX (z) if z ∈ V \ P ,  n n+1 X  Φn+1(z) = HQ(z) if z ∈ Q ∈ Pn+1 and Q ⊂ Yn+1,  0 ϕQ(z) if z ∈ Q ∈ Pn+1 and Q 6⊂ Yn+1.

X 00 By the Gluing Lemma (Lemma 8.2), the map Φn : V → Ve is a K -quasiconformal X homeomorphism for each n > 0. Note that the sequence (Φn ) eventually stabilizes on T 0 T V\ n Yn. The set E(Y ) := n Yn consists of points in V which are never mapped 0 0 0 into Y , and hence E(Y ) ⊂ Koff-crit(F ). Since F is dynamically natural, E(Y ) has zero 0 X Lebesgue measure. Hence V\ E(Y ) is a dense set. We conclude that the sequence (Φn ) converges the limiting K00-quasiconformal homeomorphism ΦX : V → Ve. This map depends on the depth of X . Letting this depth go to ∞, in the limit we obtain the required quasiconformal map Φ: V → Ve with quasiconformal dilatation equal to max(K,K0). For this homeomorphism the properties (1), (2) and (4) follow directly from the construction, while property (3) follows from the facts that meas(E(Y0)) = 0 and that H conjugates F S and Fe on c∈S fib(c).  8.4. The QC-Criterion. The following criterion for the existence of quasiconformal ex- tensions was proven in [KSS1, Appendix 1]. This result is inspired by the works of Heinonen–Koskela [HK] and Smania [Sm3]. Recall that a topological disk P ⊂ C has η-bounded geometry if there is a point x ∈ P such that P contains an open round disk of radius η · diam(P ) centered at x.

Theorem 8.6 (QC-Criterion). For any constants 0 6 k < 1 and η > 0 there exists a constant K with the following property. Let ϕ:Ω → Ωe be a quasiconformal homeomorphism THE DYNAMICS OF COMPLEX BOX MAPPINGS 59 between two Jordan domains. Let X be a subset of Ω consisting of pairwise disjoint open topological disks. Assume that the following hold: (1) If P is a component of X, then both of P and ϕ(P ) have η-bounded geometry, and moreover mod(Ω \ cl (P )) > η, mod(Ωe \ cl (ϕ(P ))) > η; ¯ (2) |∂ϕ| 6 k|∂ϕ| holds almost everywhere on Ω \ X. Then there exists a K-quasiconformal map ψ :Ω → Ωe such that ψ = ϕ on ∂Ω.  The QC-Criterion will be applied in situations where Ω is a puzzle piece and X is a union of first return domains to Ω. Note that a priori the set X can be pretty “wild”, see Figure 15.

...... Ω ......

......

... X ...

......

Figure 15. The set X in the statement of the QC-Criterion can have a complicated shape. In particular, it is possible that X tiles Ω, as shown on the picture (components of X are shown in yellow). In this example, the second condition in the QC-Criterion is automatically satisfied: the set Ω \ X, shown in green, has zero area.

9. Sketch of the proof of the QC-Rigidity Theorem In this section we will explain how to combine shrinking of puzzle pieces (Theorem 6.1 (1)), complex bounds (Theorem 7.1), the QC-Criterion (Theorem 8.6) and the Spreading Principle (Theorem 8.5) to prove qc rigidity for non-renormalizable complex box mappings. The argument follows that of [KSS1, Section 6]. Complex bounds from Theorem 7.1 imply the following lemma, which gives us that for any point z, the orbit of which accumulates on a critical fiber, there are arbitrarily deep THE DYNAMICS OF COMPLEX BOX MAPPINGS 60 puzzle pieces containing z that satisfy the geometric hypotheses of the QC-Criterion. The details are verbatim as in [KSS1, Corollary 6.3] (simplified in several places as we do not have to deal with (eventually) renormalizable critical points). Lemma 9.1. [KSS1, Corollary 6.3 (Geometric control for landing domains)] Let F be a non-renormalizable dynamically natural complex box mapping. Then for every integer n > 0 there exists a nice puzzle neighborhood W of Critac(F ) such that for every component U of L(W ) the following hold: • U has η-bounded geometry; • the depth of U is larger than n, and if P is the puzzle piece of depth n containing U, then mod(P \ U) > η, where η > 0 is a constant independent of n. Furthermore, if Fe is another non-renormalizable dynamically natural box mapping com- binatorially equivalent to F , then taking η smaller if necessary, the same claims as above hold for Fe and the corresponding objects with tilde.  The following lemma implies that we can modify a combinatorial equivalence H from Theorem 6.1(3) so that it become a conjugacy on the orbits of critical points in Crit( F ) \ Critac(F ) (if this set is not empty).

Lemma 9.2 (Making H partial conjugacy). Let F : U → V and Fe : Ue → Ve be non- renormalizable dynamically natural box mappings that are combinatorially equivalent with respect to some K-quasiconformal homeomorphism H : V → Ve that satisfies the assump- 0 tions of Theorem 6.1 (3). Suppose Crit(F ) \ Critac(F ) 6= ∅. Then there exists a K - quasiconformal homeomorphism H0 : V → Ve so that 0 • H |∂U = H|∂U ; • H0 is a combinatorial equivalence between F and Fe; • Fe ◦ H0 = H0 ◦ F on ∂U; 0  i • H conjugates F to Fe on F (c): c ∈ Crit(F ) \ Critac(F ), i > 0 .

Proof. Recall that S := Crit(F ) \ Critac(F ) consists of non-recurrent critical points whose orbits do not accumulate on other critical points (we can talk about points rather than fibers since by Theorem 7.1(1) all critical fibers are trivial). Therefore, we can find a sufficiently deep puzzle neighborhood U0 of Crit(F ) consisting of puzzle pieces of the same ˆ depth, say n0 > 0, so that the orbit of F (S) is disjoint from U0, and hence from L(U0). For simplicity we assume that F (S) contains no critical points; otherwise, the modification of the argument below is straightforward. By increasing n0 we can also assume that each component of U0 contains only one critical point. 0 Let ϕ: U0 → Ue0 be a K -quasiconformal map that respects the boundary marking in- 0 duced by H constructed by Lemma 8.4. Clearly, K > K and can be arbitrary large. Let 0 H0 : V → Ve be the K -qc extension of ϕ by the Spreading Principle (Theorem 8.5). If P is a sufficiently small puzzle piece around a point x ∈ orb(F (S)), then P is not contained ˆ in L(U0) by our choices, and thus by property (4) of the Spreading Principle, H0(P ) = Pe. THE DYNAMICS OF COMPLEX BOX MAPPINGS 61

Since puzzle pieces shrink to points (Theorem 6.1(1)), by letting the depth of such P ’s go k k to zero we conclude that H0(F (c)) = Fe (ec) for every c ∈ S and k > 1. 0 0 −1 Now define H : V → Ve by H = Fe ◦ H0 ◦ F on the components of U0 containing the 0 k k points of S and H = H0 elsewhere. As H0(F (c)) = Fe (ec) for all k > 1 and c ∈ S, the 0 map H is well-defined and it is continuous because Fe ◦ H0 = H0 ◦ F on ∂U0. It follows 0 0 that H is the required K -qc map. 

Lemmas 9.2, 9.1 and the QC-Criterion (Theorem 8.6) imply the following proposition (which is similar to the claim in [KSS1, Section 6.4]). The conclusion of this proposition will be then an input into the Spreading Principle later in the proof of qc rigidity of complex box mappings.

Proposition 9.3 (Uniform control of dilatation). Let F and Fe be non-renormalizable dynamically natural box mappings that are combinatorially equivalent with respect to some homeomorphism H that satisfies the assumptions of Theorem 6.1 (3). Then there exists K > 0 such that for every critical point c ∈ Crit(F ) and the cor- responding critical point ec ∈ Crit(Fe), and every n > 0 there exists a K-quasiconformal homeomorphism respecting the boundary marking between the corresponding critical puzzle pieces of depth n containing c and ec. Proof. Suppose F and Fe are combinatorially equivalent with respect to a k-quasiconformal homeomorphism H. By Lemma 9.2, we can assume that H is a conjugacy between F and Fe on the orbits of points in Crit(F ) \ Critac(F ). Let Q, Qe be the pair of corresponding puzzle pieces of depth n containing c, respectively ec. If c ∈ Crit(F ) \ Critac(F ), then the required homeomorphism between Q and Qe is obtained just by pulling back H. So assume that c ∈ Critac(F ). For the chosen n, let W and Wf be a pair of neighborhoods of Critac(F ) and Critac(Fe) given by Lemma 9.1, and let ϕ0 : W → Wf be a K0-quasiconformal map that respects the boundary marking given by Lemma 8.4. Note that K0 can be arbitrarily large. We can spread this map around using the Spreading Principle (Theorem 8.5). Restricting the resulting map to the depth n puzzle piece Q we obtain a max(k, K0)-quasiconformal map ϕ: Q → Qe that respects the boundary marking. Let X be the union of the landing domains to W that are contained in Q. By Theo- rem 8.5, the map ϕ has the property that its dilatation is bounded by k on Q \ X. On the other hand, each component P of X satisfies the conclusion of Lemma 9.1, that is, P has η-bounded geometry and mod(Q \ P ) > η for some constant η independent of n. The same is true for Qe and Pe (with the same constants). Therefore, we can apply the QC-Criterion to Ω := P ⊃ X, Ωe := Pe ⊃ Xe and the map ϕ. This criterion will give us a desired K-quasiconformal homeomorphism between Q and Qe that agrees withϕ ˆ on ∂Q, and hence respects the boundary marking. The proof is complete once we note that by the QC-Criterion, the dilatation K depends only on k and η, but not on n, or on K0. Further details of the proof can be found in [KSS1, Page 787].  THE DYNAMICS OF COMPLEX BOX MAPPINGS 62

Finally, we are ready to present the proof of qc rigidity for non-renormalizable box mappings. Proof of Theorem 6.1 (3). Suppose that the homeomorphism H that provides the com- binatorial equivalence between F and Fe is k-quasiconformal. For every n > 0, define S Wn := c∈Crit(F ) Pn(c), where Pn(c) is the puzzle piece of depth n containing c, and simi- larly define Wfn for Fe. By Proposition 9.3, we can construct a sequence of K-quasiconformal maps hn : Wn → Wfn so that each of them respects the boundary marking induced by H. Since Crit(F ) ⊂ Wn and hn respects the boundary marking, we can spread hn around using Theorem 8.5. In this way we obtain a sequence of max(k, K)-quasiconformal maps Hn : V → Ve each conjugating F and Fe outside of Wn. Finally, by Theorem 7.1(1), the sets Wn shrink to Crit(F ) as n → ∞. Therefore, by passing to a subsequence, Hn converges to a quasiconformal map H∞ : V → Ve that conjugates F and Fe. Using the properties of Hn from Theorem 8.5, it is not hard to see that H∞ does not depend on the subsequence. This concludes the proof.  10. Sketch of the proof of Complex Bounds As was discussed in Section 5.2, there are three combinatorially distinct types of criti- cal points: non-combinatorially recurrent, reluctantly recurrent and persistently recurrent. The proofs of complex bounds around each of these types of critical points are different from one another. For reluctantly recurrent and non-recurrent critical points the claim of Theorem 7.1 follows from the fact that around each of these points one can find arbitrarily deep puzzle pieces which map with uniformly bounded degree to some piece with bounded depth. For persistently recurrent critical points the proof will employ the Enhanced Nest construction of Section 7.4. These arguments will be made more precise in Lemmas 10.1, 10.2, and 10.3 below, dealing with reluctantly, non-, and persistently recurrent cases re- spectively. To obtain Theorem 7.1 from these lemmas, which give, in particular, complex bounds for induced complex box mappings around recurrent critical points, a little care is needed, since critical fibers can accumulate on another critical fibers. However, this step is straight- forward and is done verbatim as in the proof of [KSS1, Proposition 6.2], and hence will be omitted in our exposition. Our focus will be only on those lemmas. 10.1. Puzzle control around non- and reluctantly recurrent critical points. First we discuss puzzle pieces around a reluctantly recurrent point c. Lemma 10.1 (Reluctantly recurrent case). Let c be a reluctantly recurrent critical point. Then there exists a positive constant η > 0 so that for any  > 0 there are nice puzzle neighborhoods W 0 ⊃ W of Back(c) such that: (1) Each component of W has η-bounded geometry. 0 0 0 (2) For each c ∈ Back(c), we have diam(Wc0 ) 6  and mod(Wc0 \ W c0 ) > η; here 0 0 Wc0 = Compc0 W and Wc0 = Compc0 W . k 0 0 (3) F (∂Wc0 ) ∩ Wc = ∅ for each c ∈ Back(c) and each k > 1. THE DYNAMICS OF COMPLEX BOX MAPPINGS 63

Proof. Lemma 6.5 in [KSS1] uses the definition of reluctant recurrence to show that one can map deep critical pieces around a reluctantly recurrent critical point to one of arbitrary fixed depth with uniformly bounded degree. Then Lemma 6.6 in that paper shows that one obtains puzzle pieces with bounded geometry and moduli bounds around c0 ∈ Back(c). The proofs use only condition (1) of Definition 4.1 (in the form of Lemma 4.2) and goes verbatim as in the proof of [KSS1, Lemma 6.6].  Moving on to non-recurrent critical points, we have the following statement. Lemma 10.2 (Non-recurrent case). Let c be a non-recurrent critical point. Then there is a constant η > 0 and for every  > 0 there exists a puzzle piece W 3 c such that diam(W ) 6  and, if c ∈ Critac(F ), such that W has η-bounded geometry. Proof. As in [KSS1, page 782], we only need to consider the case when the orbit of c does not intersect or accumulate on a reluctantly recurrent critical point. We then distinguish three cases, depending whether Forw(c) is 1) empty; 2) not empty, and contains only combinatorially non-recurrent critical points; and 3) not empty and contains a persistently recurrent critical point. (Case 4 in [KSS1] concerns renormalizable critical points which we do not have in our setting by hypothesis.) Note that case 1) corresponds to the situation when c 6∈ Critac(F ). k In Case 1, choose n0 to be the depth of puzzle pieces such that Pn0 (F (c)) contains no critical points for every k > 1 (here Pk(x) stands for the puzzle piece of depth k containing the point x). Since F is dynamically natural, condition (3) of Definition 4.1 guarantees K(F ) = Kwell-inside(F ). Therefore, there exists δ > 0 such that the orbit of c visits infinitely many times the components of U that are δ-well inside the respective components of V. n−1 n Hence there exist infinitely many n > 1 such that F : Pn0+n−1(F (c)) → Pn0 (F (c)) is a conformal map of uniformly bounded distortion; this distortion depends on δ. The conclusion in case 1) follows. (Note that in this case we do not claim that W has bounded geometry.) Case 2: Since Forw(c) does not contain combinatorially recurrent critical points, we 0 0 0 can fix n0 so that for each c ∈ Forw(c) the forward orbit of c is disjoint from Pn0 (c ) k 00 00 00 and so that if F (c) ∈ Pn0 (c ) for some c ∈ Crit(F ) and k > 0, then c ∈ Forw(c). 0 0 0 0 Choose c ∈ Forw(c). Fix n > n > n0 so that Pn0 (c ) b Pn(c ). Let ni be a se- n 0 0 i 0 quence of times so that F (c) ∈ Pn (c ) and let Pni+n(c) be the pullback of Pn(c ) under ni ni 0 the map F . We claim that F : Pni+n(c) → Pn(c ) has bounded degree. Indeed, if k c1 ∈ F (Pni+n(c)) for some 0 6 k < ni and for some c1 ∈ Crit(F ), then c1 ∈ Forw(c) k and, moreover, since F (Pni+n(c)) ⊂ Pn0 (c1) no further iterate of Pni+n(c) can con- tain c1. It follows that each critical point in Crit(F ) can be visited at most once by n 0 0 0 i 0 Pni+n(c),F (Pni+n(c)),...,F (Pni+n(c)) = Pn(c ). Note that Pn (c ) has some η -bounded geometry (where η0 depends of course on the integer n0 which we fixed once and for all). 0 0 n 0 0 i Since Pn (c ) b Pn(c ) and the degree of F : Pni+n(c) → Pn(c ) is bounded in terms of the number and degrees of the critical points of F , it follows that Pni+n(c) has η-bounded geometry, where η does not depend on i. By Montel (or Gr¨otzsch) it then also follows that the diameters of Pni+n(c) must shrink to zero. THE DYNAMICS OF COMPLEX BOX MAPPINGS 64

The proof in Case 3 goes verbatim as in the proof of [KSS1, Lemma 6.8], where in fact the set Ωi in that paper can be replaced by a puzzle piece in our setting. (In fact, this would allow us to simplify the argument in [KSS1] for this case, because there the set Ωi is obtained via the upper and lower bounds in [KSS1] which are proved by hand, whereas here we can use the Enhanced Nest construction combined with the Covering Lemma.)  10.2. Puzzle control around persistently recurrent critical points. Lemma 10.3 (Persistently recurrent case). Let c be a persistently recurrent critical point. Then there exists δ > 0 so that for each  > 0 there exists a nice puzzle neighborhood W of [c] with the following properties: (1) Each component of W has δ-bounded geometry. 0 ˆ (2) For each c ∈ Back(c), we have diam(Wc0 ) 6 , where Wc0 = Lc0 (W ). (3) The first landing map to W is (δ, N)-extendible w.r.t. [c]. Proof. The proof of this result the same as the proof of [KSS1, Lemma 6.7]. It follows by inducing a persistently recurrent complex box mapping from F , see Lemma 7.5, and the Key Lemma (Lemma 10.4) below.  Finally, let us state and outline the proof of the Key Lemma. Because of Lemma 7.5, we can assume that our box mapping is persistently recurrent by considering an induced map. Lemma 10.4 (Key Lemma). Let F : U → V be a persistently recurrent non-renormalizable dynamically natural box mapping. Then there exists η > 0 so that for every c ∈ Crit(F ) and every  > 0 there is a pair of puzzle pieces Y 0 ⊃ Y 3 c such that (1) diam Y < ; (2) Y has η-bounded geometry at c; 0 0 (3) (Y \ Y ) ∩ PC(F ) = ∅, and mod(Y \ Y ) > η. Proof. This lemma was proven in Sections 10 and 11 of [KvS] using the Enhanced Nest construction presented in Section 7.4 and the Covering Lemma of Kahn and Lyubich (Lemma 7.10). Let us sketch the argument.

Let c0 ∈ Crit(F ) be a critical point, necessarily persistently recurrent, and let (In)n>0 be the Enhanced Nest around c0. Recall that a puzzle piece P around c0 is called δ-nice if mod(P \ Compx R(P )) > δ for every x ∈ P ∩ PC(F ). Note that since F is persistently recurrent, there are only finitely many components in R(P ) intersecting PC(F ). The key step in [KvS] was to prove:

12 0 Proposition 10.5. [KvS, Proposition 10.1] There exists a beau δ > 0 and N ∈ N, 0 (depending on F ), so that for all n > N, In is δ -nice. In particular, for every  > 0, there exists δ > 0, so that if I0 is -nice, then In is δ-nice for every n > 1.

12That is, a constant depending only on the number and degrees of the critical points of F . THE DYNAMICS OF COMPLEX BOX MAPPINGS 65

We will prove this proposition after we show how it implies Lemma 10.4. This proposition uses for its proof, in a non-trivial way, the results of Lemma 7.8 on transition and return times for the Enhanced Nest and the Covering Lemma (Lemma 7.10). We will sketch its proof below; for now, let us use this proposition and complete the proof of the Key Lemma. Since each In+1 is a pullback of In containing c0, Proposition 10.5 implies that mod(In \ In+1) > δ. Hence, by the Gr¨otzsch inequality, the puzzle pieces In, and thus all puzzle pieces around c0 shrink to zero in diameter. ν By the discussion after Lemma 7.6 , the degree of the map F |B(In) does not depend on n, and so Proposition 10.5 gives us that the moduli mod(B(In) \ A(In)) are uniformly bounded from below (see also Figure 12, left). Since the mapping from In+1 to In has − + uniformly bounded degree, one can construct puzzle pieces In+1 ⊂ In+1 ⊂ In+1 so that + − mod(In+1 \ In+1) and mod(In+1 \ In+1) are both uniformly bounded from below, and disjoint from PC(F ). In other words, the puzzle pieces In are all δ-free (see Section 1.5.5 for the definition). This property together with the Koebe Distortion Theorem implies 0 0 that if In has δ-bounded geometry at the critical point, then F (In+1) has δ = δ (δ) bounded geometry at the critical value, and pulling back F (In+1) by one iterate of F to In+1 improves the bounded geometry constant. See [KvS, Proposition 11.1] for details. Now the Key Lemma for arbitrary c ∈ Crit(F ) follows by picking Y to be Lˆc(In) and 0 ˆ + Y to be Lc(In ) for n sufficiently large.  Let us finally sketch the proof of Proposition 10.5. Proof of Proposition 10.5. We will use the results and notation from Section 7.4. Pick some big integer M and assume that n > M is sufficiently large. Let A be a component of R(In) intersecting PC(F ). Our goal is to estimate mod(In \ A) and show that for all sufficiently large n such moduli are uniformly bounded from below. t Define tn,M := pn−1 + pn−2 + ··· + pn−M and notice that F n,M (In) = In−M . It fol- lows from Lemma 7.8 (see also the discussion at the end of Section 7.4) that F tn,M (A) is contained in some landing domain to In−4. t Choose any z ∈ F n,M (A) ⊂ Lz(In−4), and any integer k satisfying n − 4 6 k 6 n − M. −νk + Let νk > 0 be the entry time of z into Ik, and set Ak be the component of F (Ik \ Ik) + that surrounds z (where the outer puzzle piece In was defined before Lemma 7.8). Thus t we obtain M − 6 annuli Ak that are contained in In−M , and all of them surround F n,M (A) −t (see Figure 16). The components of F n,M (Ak) which surround A give us M − 6 annuli contained in In which surround A. Using the Covering Lemma, we can show that by taking M sufficiently large, the sum of their moduli is bounded. This yields a lower bound on mod(In \ A). Let us explain how this is done.

Let µn > 0 be so that In is µn-nice. When F is persistently recurrent, the set PC(F ) is + compact, so µn is not zero. The construction of the Enhanced Nest gives us annuli Ik \ Ik that are disjoint from PC(F ) and with moduli bounded from below by K1µk−1, where K1 ν depends only on the degrees of the critical points of F . Since F k | + is a landing map, Lz(Ik ) THE DYNAMICS OF COMPLEX BOX MAPPINGS 66

q2341

Figure 16. The construction of pullback annuli in the proof of Proposi- tion 10.5.

its degree is bounded in terms of the degrees of the critical points of F , and so there is a constant K with the property that mod(A ) K1 µ . This gives us a lower bound on 2 k > K2 k−1 t the moduli of the annuli Ak surrounding F n,M (A). Now we are going to use the Covering Lemma (Lemma 7.10) to transfer this bound to the annuli surrounding A. To do this, we t need to bound the degree of F n,M |A independently of n and M, but this is not too hard, and is done in Step 5 of the proof in [KvS, Proposition 10.1], with the key idea essentially presented at the end of Section 7.4. Let d denote the bound on the degree of this mapping. Finally we will explain how the Covering Lemma is applied. Let µn−5,M−5 = min(µn−5, µn−6, . . . , µn−M ). Taking K1 smaller if necessary, we may assume that K1/K2 < 1. We let C > 0 be the universal constant from the Covering Lemma. tn,M Fix η = K1/2K2 and D to be degree of F |In . Let  > 0 be the constant associated 0 to η and D by the Covering Lemma. For each annulus Ak, let B, respectively B , be the l 2 m regions bounded by the inner, respectively outer, boundaries of A . Fix M = 4d K2 + 6 , k CK1η and let V = In−M . Then we have that 2 K1 4d mod(V \ B) > (µn−M−1 + ··· + µn−5) > µn−5,M−5. K2 Cη Thus the Covering Lemma implies that −1 −2 (10.1) mod(In \ A) > min(, η mod(Ak), Cηd mod(V \ B)) > min(, 2µn−5,M−5).

We always have that there exists a constant K so that µn > Kµn−1. Combining these estimates it is easy to argue that the µn’s stay away from 0. Finally, to see that there exists δ0 > 0, beau, as in the statement of the proposition, observe that by (10.1) (because of the multiplication by 2 of µn−5,M−5) this bound is beau provided that  does not depend on F . The constant  = (η, D) is given by the Covering Lemma, and we have that η = η(K1,K2) and D = D(M) = D(d, K1,K2,C). Notice that THE DYNAMICS OF COMPLEX BOX MAPPINGS 67 the constants d, K1,K2 are universal constants depending only on the number and degrees of the critical points of F and C is a universal constant given by the Covering Lemma. 

11. All puzzle pieces shrink in diameter In this section, we show how to conclude Theorem 6.1(1) using Theorem 7.1. For critical points the claim is established in Theorem 7.1(1). Our goal here is to show shrinking of puzzle pieces around all other points in the non-escaping set K(F ) of F . Let x ∈ K(F ) \ Crit(F ) be a point. We consider two cases: (1) the orbit of x accumulates at some critical fiber fib(c); (2) the orbit of x is disjoint from some puzzle neighborhood of Crit(F ). We start with the first case. Let  > 0, and let W be the nice puzzle neighborhood of Crit(F ) given by Theorem 7.1 with the property that each of its components has Euclidean diameter at most . Since orb(x) accumulates at fib(c), for some c ∈ Crit(F ), there exists s s s ∈ N so that F (x) ∈ W. We let W denote the component of W that contains F (x), and we let U = Lc(W ). By Theorem 7.1(3), the corresponding branch of the first landing map is (δ, N)-extendible with respect to Crit(F ), where neither δ, nor N depends on . So there exists an arbitrarily small puzzle piece Wf c W, such that mod(Wf \ W ) > δ, and −s 0 0 0 if Ue = Compx F (Wf), we have that mod(Ue \ U) > δ , where δ = δ (δ, N). Thus by the Gr¨otzsch inequality we have that the puzzle pieces around x shrink to points. Let us treat the second case. Since F is dynamically natural, K(F ) = Kwell-inside(F ) (condition (3) of Definition 4.1). Therefore there exist an infinite sequence of components j j j j j j U of U and V of V and a δ = δ(x) > 0 such that U ⊂ V , mod(V \U ) > δ, and the orbit S j of x visits X := j U infinitely often. Choose m sufficiently large so that orb(x) does not intersect the puzzle neighborhood of Crit(F ) of depth m, and suppose m < m1 < m2 < . . . mi mi is a sequence of iterates such that F (x) ∈ X, and let U = CompF mi (x) X. Let Ui, Vi be the pair of puzzle piece, both containing x, obtained by pulling back U mi and V mi under m F i . By construction, each Vi \ U i is a non-degenerate annulus. Moreover, the degree of m m the map F i : Vi → V i is bounded above by some constant C > 1 that depends only on m and the local degrees of F at the critical points, but not on k. Hence, by the Koebe 0 0 0 Distortion Theorem, there exists δ = δ (δ, C) > 0 such that mod(Vi \ U i) > δ for every i > 0. Now the conclusion follows by the Gr¨otzsch inequality once we arrange the sequence (mi) so that Vi+1 ⊂ Ui for every i > 0. 

12. Ergodicity properties and invariant line fields In this section, we present the proof of Theorem 6.1(2a). Since this theorem is auto- matically true if Crit(F ) = ∅ (in this case meas(K(F )) = 0 by definition of dynamical naturality), in this section we assume that F has at least one critical point. By Theorem 6.1(1), the puzzle pieces of a non-renormalizable dynamically natural box mapping F shrink to points. Therefore, the orbit of a point z ∈ K(F ) accumulates on the fiber of a point z0 ∈ K(F ) (i.e. orb(z) enters every small puzzle neighborhood of z0) if and only if z0 ∈ ω(z). Furthermore, for a pair of distinct points z, z0 ∈ K(F ), every two nests THE DYNAMICS OF COMPLEX BOX MAPPINGS 68 of puzzle pieces around them are eventually disjoint (starting from some large depth). We will frequently use these two observations below without further notice.

12.1. Ergodicity properties. In this subsection, we study ergodic properties of non- renormalizable dynamically natural box mappings. In particular, in the next theorem we draw consequences for the map in presence of a forward invariant set of positive Lebesgue measure. The real analogue of this result was proved in [SV], see also [McM, Theorem 3.9] and [Man]. Even though this result is interesting in its own right, our purpose is to use it in Section 12.2 in the situation when this forward invariant set contains the support of a measurable invariant line field.

Theorem 12.1 (Ergodic properties of non-renormalizable box mappings). Let F : U → V be a non-renormalizable dynamically natural box mapping with Crit(F ) 6= ∅, and let X ⊆ K(F ) be a forward invariant set of positive Lebesgue measure. Let z be a Lebesgue density point of X. Then the orbit of z accumulates on a critical point c, and any such critical point c is a point of Lebesgue density of X. Moreover, if c is not persistently recurrent, then there exist puzzle pieces V ⊃ K c J 3 c kn of F with meas(J ∩ X) = meas(J ) and an infinite sequence (kn) such that F (z) ∈ J kn −kn and the maps F : Compz F (K) → K have uniformly bounded degrees (independent of n).

Proof. Since F is dynamically natural, meas(Koff-crit(F )) = 0, and hence the orbit of Lebesgue almost every point in X accumulates on a critical point. Pick z ∈ X to be a Lebesgue density point of X. We can assume that z is not a backward iterate of a critical point, and hence focus only on critical points in ω(z). Define a partial ordering on the set of critical points of F as follows. Given two critical 0 0 k 0 0 points c, c ∈ Crit(F ), we say that c > c if F (c) = c for some k > 0, or ω(c) 3 c . 0 0 Note that c > c implies ω(c) ⊇ ω(c ). Among all critical points in ω(z), let c0 ∈ ω(z) be a maximal critical point with respect to this partial ordering. Note that there might be several such points, we pick one of them. Recall from Section 5.2,[c0] stands for the equivalence class of critical points c such that c ∈ ω(c0) and c0 ∈ ω(c) (as usual, the discussion from Section 5.2 simplifies since we know that the fibers of F are all trivial). Let c0 > c1 > c2 > ... > ck be a chain in the partial ordering starting from c0 and consisting of pairwise non-equivalent critical points (we pick a representative in each equivalence class13, and the proof below repeats for each such choice). It follows from the definition of persistent recurrence that the omega-limit set of a persistently recurrent critical point c is minimal (see Section 1.5), provided that puzzle pieces shrink to points. In particular, if z ∈ ω(c) and c is persistently recurrent, then

13 A more precise way would be to write [c0] > [c1] > ... > [ck]. THE DYNAMICS OF COMPLEX BOX MAPPINGS 69

ω(z) = ω(c). Therefore, ck is the only critical point that can be persistently recurrent in the chain above14. Case 1: c0 is persistently recurrent. In this case, k = 0 and the chain consists of one element (the equivalence class [c0]). It follows that there exists an integer n0 so that if k some iterate of z, say F (z), is contained in a puzzle piece P 3 c0 of depth at least n0, then any critical point c ∈ Crit(F ) which is contained in a pullback of P along (12.1) k {z, . . . , F (z)} satisfies c ∈ [c0].

Since c0 is persistently recurrent, one can construct an Enhanced Nest (In) around c0, see Section 7.4 This nest has the property that diam In → 0 as n → ∞ and that there exists − − δ > 0 so that for every n > 0 there exists a puzzle piece In ⊂ In such that In ∩ ω(c0) ⊂ In − and mod(In \ In ) > δ. The latter bound is Proposition 10.5. Let us further assume that this nest is built sufficiently deep so that (12.1) holds for each of its elements. Note here that a priori some critical orbits for critical points not in [c0] might intersect the annulus − − k − In \In . However, due to (12.1), the inclusion In ∩ω(c0) ⊂ In implies that if F (z) ∈ In for k −k − − some k > 1, then the map % := F : Compz F (In ) → In has an extension to a branched covering onto In of the same degree as %. Taking k minimal and using this extension, by a standard argument we conclude that c0 is a Lebesgue density point of X. Case 2: c0 is reluctantly recurrent. From the definition of reluctant recurrence it follows that there exists a shrinking nest of puzzle pieces (Qi)i>0 around c0 and an increasing s s sequence of integers si → ∞ so that F i (Qi) = Q0 and so that all the maps F i : Qi → Q0 15 have uniformly bounded degrees (see [KSS1, Lemma 6.5] ). Let K = Q0 and J = Qi0 0 with the smallest i > 0 so that Qi0 b Q0 (which exists as the nest is shrinking). 0 ` Finally, for i > i , let `i be the first entry time of the orbit of z to Qi. Then F i (z) ∈ `i −`i Qi ⊂ J and the degree of the map F : Compz F (Qi) → Qi is bounded independently `i+si −(`i+si) of i. Hence, the maps F : Compz F (Q0) → Q0 = K have uniformly bounded 0 degrees for all i > i . Using a standard argument we conclude that c0 is a Lebesgue density point of X and moreover meas(J ∩ X) = meas(J ). 0 Now, for j ∈ {1, . . . , k}, let J be an arbitrary small puzzle piece around cj of depth s 0 larger than the depth of J . Suppose s > 0 is the smallest iterate so that F (c0) ∈ J . Then −s 0 −s 0 meas(X ∩ Compc0 F (J )) = meas(Compc0 F (J )). Since F is a non-singular map and X is forward invariant, we can iterate X forward to obtain meas(J 0 ∩ X) = meas(J 0). 0 As this is true for arbitrary small puzzle pieces J 3 cj, we conclude that cj is a point of Lebesgue density for X. Case 3: c0 is non-recurrent. The proof in this case is similar to the proof in Case 2. First observe that [c0] = c0 and no other critical point in ω(z) can accumulate on c0. Therefore, there exists a puzzle piece K 3 c0 of sufficiently large depth so that the orbits of points in ω(z) ∩ Crit(F ) are disjoint from K. Hence, if J 3 c0 is another puzzle piece

14This justifies the term “minimality” for the omega-limit set of persistently recurrent critical points: such points can be only at the end of any chain, i.e. persistently recurrent points are always minimal with respect to our order. 15The proof of this lemma goes through verbatim in our setting. THE DYNAMICS OF COMPLEX BOX MAPPINGS 70

k so that J b K (again, such exists since fib(c0) = {c0}) and k is such that F (z) ∈ J , then the map F k : Comp F −k(J ) → J is univalent and extends to a univalent map over K. The proof now goes exactly as in Case 2.  As usual, an ergodic component of F is a set E ⊂ K(F ) of positive Lebesgue measure so that F −1(E) = E up to a Lebesgue measure zero set. The corollary below shows that there are only finitely many of such sets that are disjoint up to a set of zero Lebesgue measure. Corollary 12.2 (Number of ergodic components). If F : U → V be a non-renormalizable dynamically natural complex box mapping, then for each ergodic component E of F there exist one or more critical points c of F so that c is a Lebesgue density point of E. 12.2. Line fields. Recall that for a holomorphic map f and a set B in the domain of f, a line field on B is the assignment µ(z) of a real line through each point z in a subset E ⊂ B of positive Lebesgue measure so that the slope of µ(z) is a measurable function of z.A −1 line field is invariant if f (E) = E and the differential Dzf transforms the line µ(z) at z to the line µ(f(z)) at f(z). The assumption that the line field is measurable, implies that Lebesgue almost every x ∈ B is a point of almost continuity: for any  > 0 meas({z ∈ D(x, r): |µ(z) − µ(x)| }) > → 0 meas(D(x, r)) as r → 0 (here D(x, r) stands for the disk of radius r centered at x). Alternatively one can define µ as a measurable Beltrami coefficient µ(z)dz/dz¯ with |µ| = 1. We say that the line field µ is holomorphic on an open set U if µ =ϕ/ ¯ |ϕ| a.e. on U, where ϕ = ϕ(z)dz2 is a holomorphic quadratic differential on U. We say that the line field is univalent if ϕ has no zeroes. In that case µ induces a foliation on U given by trajectories of ϕ. The absence of measurable line fields was proved by McMullen for infinitely renormal- izable quadratic polynomials with complex bounds [McM, Theorem 10.2], in [LvS3] for covering maps of the circle and for real rational maps with all critical points real in [She1]. Since then for a number of other cases, including for non-renormalizable polynomials [KvS]. In our exposition here we will use the following result due to Shen, [She1, Corollary 3.3] the proof of which in our setting goes through without any changes. Proposition 12.3 ([She1], Corollary 3.3). Let F be a dynamically natural complex box mapping. Let z be a point in the filled Julia set K(F ). Assume there exist a positive constant δ > 0, a positive integer N > 2, sequences {sn}, {pn}, {qn} of positive integers tending to infinity, sequences {An}, {Bn} of topological disks with the following properties (see Figure 17):

sn (i) F : An → Bn is a proper map whose degree is between 2 and N; (ii) diam(Bn) → 0 as n → ∞; pn qn − 16 (iii) F (z) ∈ An, F (z) ∈ Bn and there exists a topological disk Bn ⊂ Bn so that

 sn sn 0  qn pn+sn − F (u): u ∈ An, (F ) (u) = 0 ∪ F (z),F (z) ⊂ Bn

16A small typo in [She1] is corrected in here. THE DYNAMICS OF COMPLEX BOX MAPPINGS 71

and − mod(Bn \ cl Bn ) > δ. (iv) for any n the following maps are univalent:

pn −pn qn −qn F : Compz F (An) → An and F : Compz F (Bn) → Bn. Then for any F -invariant line field µ, either z is not in the support of µ, or µ is not almost continuous at z. 

Figure 17. An illustration to the statement of Proposition 12.3. The critical values of the map F sn are marked with crosses. The green annulus has modulus at least δ > 0, with δ independent of n. The branches F pn and F qn are univalent. If z is in the support of µ and µ is almost continuous at z, then by going to limits, one can extract univalent line fields on An and sn also Bn, which is impossible because the map F : An → Bn has critical points.

The proof of the above proposition follows the same idea as the original idea by Mc- pn qn Mullen: the maps F , F send a neighborhood of z diffeomorphically to An, Bn and so since µ is increasingly constant near a point of almost continuity z, by property (iv) we obtain in An,Bn what increasingly looks like continuous foliations of trajectories of the sn corresponding quadratic differentials. However, the map F : An → Bn has critical points, so this impossible. To make this argument precise one needs in particular assumption (iii). We are now in position to prove Theorem 6.1(2a) in the form of the following theorem. Theorem 12.4 (Invariant line fields). Let F : U → V be a non-renomalizable dynamically natural box mapping with Crit(F ) 6= ∅. Assume that F supports a measurable invariant line THE DYNAMICS OF COMPLEX BOX MAPPINGS 72

field on a forward invariant subset X ⊆ K(F ) of positive Lebesgue measure. Then there exists a puzzle piece J of F so that meas(X ∩ J ) = meas(J ) and so that the invariant line field extends to an invariant univalent line field on J .

Proof. The proof uses the result of Theorem 12.1. Let z ∈ X be a Lebesgue density point which is a point of almost continuity for µ. Assume that ω(z) contains a reluctantly recurrent or non-recurrent critical point, say, c0. By Theorem 12.1, there exist puzzle pieces 0 K c J 3 c of F with meas(J ∩ X) = meas(J ) and an infinite sequence (kn) such that kn kn −kn F (z) ∈ J and the maps F : Compz F (K) → K have uniformly bounded degrees, −kn say, bounded by some N independent of n. Write Un := Compz F (J ). Since z is a Lebesgue density point of X, it follows that meas(Un ∩ X)/ meas(Un) → 1 as n → ∞ (note that (Un) form a shrinking nest of puzzle pieces with uniformly bounded geometry). kn As F |Un is a composition of univalent maps with bounded distortion and at most N unicritical covering maps with a unique critical point of the form w 7→ wdi , we obtain that µ is a holomorphic line field on J . Since µ can have only finitely many singularities on J , by, if necessary, shrinking J slightly we can obtain that the line field µ is univalent on J . So assume now that ω(z) contains only persistently recurrent critical points, and let c0 ∈ ω(z) be one of them. For c0, we start by repeating the same argument as in the proof of Case 1 of Theorem 12.1. Based on that argument, let show how to construct the topological disks An,Bn and the sequences {sn}, {pn}, {qn} with the properties required by Proposition 12.3. − ˆ − tn,c − − For each c ∈ [c0], define Wn,c := Lc(In ) and let F : Wn,c → In be the corresponding − −tn,c branch of the first landing map to In . If now Wn,c := Compc F (In), then Wn,c∩ω(c0) ⊂ W − . Define W − := S W − . n,c n c∈[c0] n,c − − − Let qn be the first entry time of orb(z) to Wn , and let Bn := Wn,c1 be the component − qn qn −qn − − of Wn containing F (z). Put Bn := Wn,c1 . The map F : Compz F (Bn ) → Bn is − univalent (because of (12.1) and since Wn ⊃ [c0]). Moreover, the inclusion Wn,c1 ∩ω(c0) ⊂ − qn −qn Wn,c1 implies that this map has a univalent extension F : Compz F (Bn) → Bn with − 0 0 mod(Bn \ cl Bn ) > δ > 0, where δ depends only on δ and the number and local degrees 0 of the critical points in [c0], in particular, δ is independent of n. The latter moduli bound tn,c − − holds because the branch F 1 : Bn → In has degree bounded independently of n, and tn,c same is true for its extension F 1 : Bn → In, again due to (12.1). Now, let V − := S Lˆ (B−), and let p be the entry time of orb(z) to V −. Let A− be n c∈[c0] c n n n n − pn sn − − the component of Vn containing F (z) and F : An → Bn be the corresponding branch − of the first landing map to Bn ; as such, the degree of this branch is bounded independently −sn − of n. Let A := Comp − F (B ). By construction, A ∩ ω(c ) ⊂ A , and hence by n An n n 0 n sn sn (12.1) the map F : A → B has the same degree as F | − , and thus it is bounded n n An pn −pn independently of n, while the map F : Compz F (An) → An is univalent. Using the disks and sequences of iterates constructed above, we are in position to apply Proposition 12.3; this gives us the desired contradiction to our choice of z to be a point of almost continuity for µ.  THE DYNAMICS OF COMPLEX BOX MAPPINGS 73

13. Lattes` box mappings In this section, we construct and describe non-renormalizable dynamically natural box mappings F : U → V with non-escaping set of positive area and which support a measurable invariant line field. We will call these Latt`esbox mappings. Such a map is quite special, and should be thought of as an analogue of a Latt`esmapping on the Riemann sphere, see [Mil3].

13.1. An example of a dynamically natural Latt`esbox mapping.

Proposition 13.1 (Latt`esbox mappings exist). There exists a non-renormalizable dynam- ically natural box mapping F : U → V with meas(K(F )) > 0 and a measurable invariant line field supported on K(F ). For the example we give here (1) Crit(F ) = {c} and F 2(c) = F (c), i.e. the unique critical point of F is strictly pre-fixed; (2) V consists of two components V and V 0 such that U := U ∩ V = V contains c and such that F (c) is contained in a component U 0 of U ∩ V 0; (3) U ∩ V 0 contains infinitely many components and they tile V 0; (4) F (U 0) = V 0 and for each component U 00 of U ∩V 0 distinct from U 0 one has F (U 00) = V .

Proof. This example starts with the box mapping F2 : U2 → V2 := (−1, 1) × (−1, 1) from Theorem 3.1(2) which was based on the Sierpi´nskicarpet construction. This box mapping leaves invariant the horizontal line field on (−1, 1) × (−1, 1). Denote the component of U2 0 containing 0 by U2 . We can make sure that U2 is invariant under z 7→ −z, i.e. if U is a component of U2, then −U is also a component of U2. Moreover, it is straightforward to 0 0 modify this map so that F2(−z) = −F2(z) for z ∈ U2 and F2(−z) = F2(z) for z ∈ U2 \ U2 , and so that F2 still preserves horizontal lines. By Theorem 3.1(2), the set K(F2) has full Lebesgue measure in V2. Moreover, we have that each puzzle piece of F2 maps linearly over V2. It follows that the orbits of Lebesgue almost all points z ∈ K(F2) will enter an arbitrarily deep puzzle piece around 0. Indeed, assume by contradiction that there exists a puzzle piece P 3 0 so that the set of points the orbits of which avoid P has positive Lebesgue measure; call this set Y . Take a Lebesgue density point y of Y and a puzzle Pn 3 y of depth n > 0; note that diam Pn → 0 n as n → ∞. Since F2 : Pn → V2 is a linear map, we have meas(Pn ∩ Y )/ meas(Pn) 6 (meas(V2) − meas(P ))/ meas(V2) < 1, which gives a contradiction. Let ϕ:(−1, 1) × (−1, 1) → D be the Riemann mapping. For simplicity choose ϕ so that it preserves the real line and so that ϕ(0) = 0. By symmetry it follows that ϕ(−z) = −ϕ(z) ˆ ˆ ˆ ˆ ˆ ˆ −1 for all z ∈ V2. Set V2 := D, U2 := ϕ(U2) and define F2 : U2 → V2 by F2 = ϕ◦F2 ◦ϕ . Then ˆ ˆ F2 also has a holomorphic invariant line field on D. Moreover, F2 preserves 0, maps real ˆ ˆ 0 0 points to real points, and is univalent on each component of U2. Let 0 ∈ U2 := ϕ(U2 ) be the ˆ ˆ ˆ ˆ 0 ˆ ˆ ‘central’ component of F2. It follows that F2(−z) = −F2(z) for z ∈ U2 and F2(−z) = F2(z) ˆ ˆ 0 for z ∈ U2 \ U2 THE DYNAMICS OF COMPLEX BOX MAPPINGS 74

Finally, using Fˆ2 we will now define a box mapping F : U → V, as described in the statement of Proposition 13.1, by taking V to be a formal union of two copies of D, and U to be a formal union of D and countably many open topological disks that tile some other copy of D. The details are as follows (note that this construction is taken for simplicity, and can be easily modified so that V would embed into C). 2 0 Let Q: D → D be the quadratic map Q(z) = z . Let U = V = V = D and define 0 0 ˆ 0 ˆ 0 F : U → V by F (z) = Q(z) for z ∈ U. Moreover, define U ∩ V := Q(U2), U := Q(U2 ) and F : U 0 → V 0 by −1 F (z) = Q ◦ Fˆ2 ◦ Q (z). ˆ ˆ ˆ 0 This map is well defined because F2(−z) = −F2(z) for all z ∈ U2 and so different choices −1 of Q (z) give the same result. Since Fˆ2(0) = 0, this map is also univalent. Finally, define F :(U ∩ V 0) \ U 0 → V by −1 F (z) = Fˆ2 ◦ Q (z).

Again this map is univalent and well-defined because Fˆ2(−z) = Fˆ2(z) for each z in a non-central component of Uˆ2. The fact that F preserves an invariant line field follows from the commuting diagram in 17 Figure 18 . Indeed, there is a holomorphic line field µ on V = ϕ(V2), which is the image under ϕ of the horizontal line field on V2 = (−1, 1) × (−1, 1) and which is invariant under 0 Fˆ2. The pushforward Q∗µ of this line field µ on V = Q(V ) by Q is again a holomorphic line field which, by definition, is invariant under F : U 0 → V 0. The map F sends each 0 0 component of (U ∩ V ) \ U univalently onto V and again the line field of Q∗µ is mapped by F to the line field µ by definition. By construction, K(F ) = Kwell-inside(F ), and by the 2nd paragraph in the proof, almost every point in K(F ) accumulates onto the critical point, and hence F is dynamically natural. 

13.2. Properties of Latt`es box mappings. Latt`esbox mappings are rather special. Some of their necessary features are described in the following result, which, in particular, provides a proof to Theorem 6.1(2b). Proposition 13.2 (Properties of Latt`esbox mappings). Assume F : U → V is a Latt`es box mapping, i.e. F is a non-renormalizable dynamically natural box mapping that has an invariant line field supported on a forward invariant measurable set X ⊆ K(F ) of positive Lebesgue measure. Let J ⊂ U be the puzzle piece from Theorem 12.4. Define Y := S F i(J ) to be the union of forward orbits for all points in J and Crit0 = {c , . . . , c } i>0 0 k to be the set of critical points of F which are contained in Y . Moreover, let PC0 :=  i F (cj): i > 0, j = 0, . . . , k . Then: (1) PC0 is a finite set, i.e. each critical point in Crit0 is eventually periodic;

17The figure was drawn using T.A. Driscoll’s Schwarz–Christoffel Toolbox for MATLAB, see https: //tobydriscoll.net/project/sc-toolbox/. THE DYNAMICS OF COMPLEX BOX MAPPINGS 75

(U ∩ V 0) \ U 0 V 0 V 0

F

U 0

Q: z 7→ z2 F Q: z 7→ z2

V V

Fˆ2

ˆ 0 U2

ϕ ϕ

F2

U2 V2

Figure 18. The construction of a Latt`esbox mapping F . The red, respec- tively blue, domains are mapped by the correspondingly colored branches of the map F over V , resp. V 0. Only the first two steps in the Sierpi´nski carpet construction for F2 are shown. Some leaves of invariant foliations are shown with thin black lines. THE DYNAMICS OF COMPLEX BOX MAPPINGS 76

(2) no forward iterate of a critical point in Crit0 is eventually mapped into another point in Crit0; (3) each critical point in Crit0 is quadratic; (4) each critical point in Crit0 has real multiplier; (5) F −1 (PC0) ∩ Y = PC0 ∪ Crit0.

In the proof of this proposition we will use the following several times:

t Observation: Suppose that F has a critical point at c of order d > 2. Then there exists 2πi/d a neighborhood U of c and a conformal function S : U → C with S(c) = c,DS(c) = e and so that F t ◦ S = F t on U. Hence if there exists an F -invariant line field which is holomorphic on U, then the line field (and the foliation defined by this line field) is also invariant under the symmetry S. This implies that the invariant line field can only be univalent on U if d = 2. Moreover, the line field cannot be univalent both in U and F t(U).

Proof of Proposition 13.2. By Theorem 12.4, meas(X ∩ J ) = meas(J ) and the invariant line field in J is univalent and so describes a smooth invariant foliation on J . t Claim 1: Let t0 > 0 be minimal so that F 0 (J ) contains a critical point c0. Then c0 is eventually periodic. i Proof of Claim 1: Assume by contradiction that F (c0), i > 0, consists of infinitely many distinct points. Let t > 0 be minimal so that F t(J ) is a connected component V ˆ of V. By choosing J smaller if necessary, we can assume that t > t0. Let V be the union i of components of V visited by F (c0), with i > t − t0. Since there are only finitely many components of V and the forward orbit of c0 is infinite, at least one of these components, say Vˆ , is visited infinitely many times and we can assume that this component is tiled by 0 t0 infinitely many components of U. Let t be so that F (J ) = Vˆ . Let v1, . . . , vn be the t0 critical values of F |J . Then there exists a puzzle piece Q ⊂ Vˆ containing some iterate of c0 which is disjoint from {v1, . . . , vn} (here we use that puzzle pieces shrink in diameter t0 i t0 to zero). Then there exists a puzzle piece P ⊂ J so that F (P ) = Q 3 F (c0) and F |P i is univalent. Hence the line field is univalent in a neighborhood of F (c0). However, by i 0 the assumption in the claim and since F (c0) ∈ Q, there exists a puzzle piece P ⊂ J so that F t00 (P 0) = Q for some t00 > 0 and so that the orbit F (P 0),...,F t00 (P 0) passes through 0 t00 c0. It follows that there exists a point z ∈ P so that DF (z) = 0. Since the line field is univalent on J ⊃ P 0 and on Q this gives a contradiction to the above observation. In fact, this argument is easy to see: if ϕ is the quadratic differential on J corresponding t0 to the univalent line field µ, then the trajectories of the quadratic differentials (F |P )∗ϕ t00 and (F |P 0 )∗ϕ on Q will not match because the former will have no singularity, while the latter will have at least one singularity corresponding to an iterate of a critical value; this is a contradiction to invariance of µ. Thus we can conclude that c0 is eventually periodic, proving the claim. Claim 2: The degree of the critical point c0 is two, and no iterate of c0 meets an- other critical point. Moreover, it is impossible that there exists a univalent line field in a neighborhood of any forward iterate of c0. THE DYNAMICS OF COMPLEX BOX MAPPINGS 77

Proof of Claim 2: This immediately follows from the above Observation. Alternatively, let q be the quadratic differential whose horizontal trajectories give foliation on J and its forward iterates. We say that q has zero of order k > 0 at some point z0 if in some local k 2 coordinate system with the origin at z0 the differential q has a form w dw . Therefore, if the foliation given by q is invariant under F , then the local degree of z0 w.r.t. F must divide k + 2 [Hub2, Section 5.3]. From the discussion in the previous paragraph it follows that since there exists a univa- lent line field near c0, i.e. the quadratic differential has zero of order 0 at c0, the degree of c0 must be two. Therefore, q has a simple pole at F (c0), and thus the foliation has a 1- prong singularity at F (c0) (see Figure 18, top left, for an example of a 1-prong singularity). i i−1 Assume by contradiction that F (c0) = c1 and that none of the points F (c0),...,F (c0) are critical. This way we have an invariant foliation near c1 with 1-prong singularity at c1. But then it is impossible to map this foliation to a foliation near F (c1) so to preserve the foliation: a 1-prong singularity has no local symmetries required if one wants to preserve the trajectories. The claim follows. Claim 3: Each critical point in Y is eventually periodic. Proof of Claim 3: Assume that some iterate of J intersects a critical point 6= c0, then t take t1 > t0 minimal so that F 1 (J ) contains a critical point c1 6= c0. Then, since c0 is i eventually periodic and F (c0) 6= c1 (by Claim 2), there exists a puzzle piece J1 ⊂ J so t1 t1 that F |J1 is univalent, and so that F (J1) 3 c1. Using exactly the same argument as in Claim 1 it then follows that c1 is also eventually periodic. 0 Claim 4: The degree of each critical point in Crit = {c0, . . . , ck} is equal to two. Moreover, there exists no univalent line field on a neighborhood of any point in PC0. Proof of Claim 4: As we saw in Claims 1 and 2, there exists a univalent line field on a neighborhood Ui of ci, i = 0, . . . , k. The claim now follows similarly as in Claim 2. Claim 5: Let p0, p1, . . . , pk be the periodic points on which the critical points c0, . . . , ck eventually land (some of these periodic points might coincide). Then the multipliers at these periodic points are real. Moreover, the line field is holomorphic on the components Vi of V containing pi. Proof of Claim 5: There exists a holomorphic line field (with a unique 1-prong singu- larity) in a neighborhood of pi. This line field can only be invariant near the periodic point pi if the multiplier is real. Moreover, by iterating F one obtains that the line field near the periodic point pi extends to a holomorphic line field on Vi. Let Ui 3 pi be the component of U containing pi. Then the only singularities of this line field on Vi are at iterates of a critical point cj ∈ Uj (assuming such a critical point cj exists). Claim 6: The only singularities of the holomorphic line field on Y are in PC0. In particular, if z ∈ Y and F (z) ∈ PC0, then either z ∈ Crit0, or z ∈ PC0. Proof of Claim 6: The first statement follows from the fact that the line field on J is assumed to be univalent, and so the only way singularities of the line field on Y can be created is by passing some critical point. The 2nd statement follows: the line field has a singularity at z if and only if it has a singularity at F (z).  THE DYNAMICS OF COMPLEX BOX MAPPINGS 78

Remark. This proposition does not imply that such a box mapping F is postcritically finite. Indeed, it is possible to take two box mappings Fi : Ui → Vi where F1 has an invariant line field (as described in Section 13.1) and F2 is one without an invariant line field and with an infinite postcritical set. Then taking the disjoint union of Ui and Vi we obtain a new map F : U → V with an invariant line field. However, if F : U → V is a Latt`escomplex box mapping, then there exist U 0 ⊂ U and 0 0 0 V ⊂ V so that F |U 0 : U → V is a postcritically finite complex box mapping. 13.3. Further remarks on Latt`esbox mappings. We end this section with some dis- cussion regarding Latt`esbox mappings, e.g. as constructed in Proposition 13.1, and their properties established in Proposition 13.2. 13.3.1. M¨obiusvs. quasiconformal conjugacy in presence of Latt`esparts. The box mapping F we constructed in Proposition 13.1 is flexible: there exists a family of maps Ft depending holomorphically on t ∈ D so that F0 = F and so that Ft is qc-conjugate but not conformally conjugate to F . An explicit way to construct such maps for t ∈ (−1, 1) is to consider a −1 family of maps exactly as before, but replacing the map F2 by the map F2,t := ht ◦F2 ◦ht, where  t  h (x, y) = x, y t 1 − t (so replacing squares by rectangles). As |t| → 1, the modulus of the central rectangle inside U2 tends monotonically to ∞ and thus one can see that Ft is not conformally conjugate to F0 for t > 0. From Theorem 6.1(3) it follows that if F,G are combinatorially equivalent non-renor- malizable dynamically natural box mappings so that there exists a qc homeomorphism H which is a conjugacy on the boundary (condition (3a) of that theorem), then F and G are qc-conjugate. Suppose that, in addition, F : U → V is not a Latt`esbox mapping, and so does not have an invariant line field. Then F,G are hybrid conjugate, i.e. the qc-dilatation of the qc-conjugacy between F and G vanishes in the filled Julia set of F . (In particular, if the filled Julia set has full measure in the V, then the qc-conjugacy is conformal.) This is in some sense an analogue of the Thurston Rigidity Theorem which states that two postcritically finite topologically conjugate maps are M¨obiusconjugate. 13.3.2. Latt`esrational maps vs. Latt`esbox mappings. Recall, that a rational map on the sphere is called Latt`es if it is a degree two quotient of an affine torus endomorphism. In [Mil3], Milnor gave a complete description of all Latt`esrational maps. For Latt`es box mappings, a similar description may be impossible, but nevertheless it seems feasible to give a complete list of all Latt`esbox mappings. Since Latt`esrational maps are examples of Collet–Eckmann maps, by the result in [PRL1, PRL2] it follows that one can induce a Latt`esbox mapping from a Latt`esrational map (see also the discussion in Section 2.4.1). Finally, property (5) in Proposition 13.2 is the analogue of the property of Latt`esrational −1 maps f : Cb → Cb that f (PC(f)) = PC(f) ∪ Crit(f), where PC(f) is the set of forward iterates of Crit(f) (the postcritical set). Latt`esrational maps play a special role in the THE DYNAMICS OF COMPLEX BOX MAPPINGS 79 study of postcritically finite18 branched coverings of the sphere. A celebrated result by W. Thurston gives a criterion when a postcritically finite branched covering of the sphere is equivalent (in some precise sense) to a rational map, see [DH2]. For this criterion to work, it is necessary to assume that the branched covering has so-called hyperbolic orbifold, which roughly translates to the fact that the corresponding Thurston σ-map is contracting on the Teichm¨ullerspace of possible possible conformal structures of the sphere marked with the postcritical set. The orbifold is non-hyperbolic if and only if f −1(PC(f)) = PC(f)∪Crit(f) and # PC(f) 6 4 (see [DH2, Lemma 2], and also [Hub3, Appendix C8]); such orbifolds are called parabolic. The maps with parabolic orbifold and # PC(f) 6 3 are always equivalent to a rational map and can be completely classified. The case when # PC(f) = 4 is more interesting. The only rational maps with this property are exactly flexible Latt`esrational maps. In general, a Thurston map f with 4 postcritical points (with hyperbolic or parabolic orbifold) admits a flat structure with 4 cone singularities and can be studied using so-called subdivision rules, see [BM].

14. A Man˜e´ Theorem for complex box mappings A classical theorem of Ma˜n´e[Man,ST] states that for a rational map, a forward invariant compact set in the Julia set is either expanding, contains parabolic points or critical points, or intersects the ω-limit set of a recurrent critical point (see also [dMvS, Section III.5] for the real version of Ma˜n´e’stheorem). In this section we will show that a similar, and in some sense stronger statement also holds for box mappings for which each domain is ‘well-inside’ its range. Let F : U → V be a complex box mapping so that each domain is δ-nice. More precisely, denote by Ux the component of U that contains a point x ∈ U and assume that there exists δ > 0 so that for each x ∈ U we have that mod(Vx \ U x) > δ. 0 0 1 0 Let F0 = F , U = U and V = V and let V be the union of the components of U i+1 i+1 containing critical points. Next, for i > 0 define inductively Fi+1 : U → V to be the first return map to Vi+1 and denote by Vi+2 the union of the components of U i+1 containing critical points. If y is a forward iterate of x under various compositions of F0,...,Fi then we define t(y, x) so that y = F t(x). Theorem 14.1 (Ma˜n´e-type theorem for box mappings). Let F : U → V be a complex box mapping so that each domain is δ-nice. For each ν > 1 and each κ > 0 there exists k−1 ν λ > 1 and C > 0 so that for all k > 0 and each x so that x, . . . , F (x) ∈ U \ V , k d(F (x), ∂V) > κ one has k k | DF (x)| > Cλ . Remark. If x is in the postcritical set of F , then it is enough that V is δ-nice. Remark. Notice that this statement is more similar to the statement of the Ma˜n´etheorem for real maps than the one for rational maps. Indeed, in the above theorem it is not

18A map f is called postcritically finite if each critical point of f is eventually periodic. In the setting of branched coverings, a critical point is a point at which the map is not locally injective. Postcritically finite topological branched coverings of the sphere are sometimes called Thurston maps. THE DYNAMICS OF COMPLEX BOX MAPPINGS 80 required that the distance of the orbit of x to the ω-limit set of any recurrent critical point is bounded away from zero. For general rational maps (rather than box mappings as above) k k the inequality | DF (x)| > Cλ fails: for example if x is in the boundary of a Siegel disk and this boundary does not containing critical points, then this inequality obviously fails. Remark. λ > 1 and C > 0 also depend on F . In the proof below we will assume for simplicity that each component of U is compactly contained in V. If this is not the case, then the proof below gives k k0 | DF (x)| > Cλ 0 j where k is the number of j’s with 0 6 j < k so that F (x) is contained in a domain which is not compactly contained in its range. Since k0 is of the same order as k, the original statement then follows. To prove Theorem 14.1 we need to first introduce some notation, and prove some prepara- tory lemmas. Denote by ||v||O be the Poincar´enorm of a vector v ∈ TxC on a simply connected open domain O 3 x and by ||v|| the Euclidean norm of v. Lemma 14.2. Let F : U → V be a complex box mapping so that each domain is δ-nice and define Fi as above. Then for each i > 0 there exist δi > 0 and τi > 1 and for each κ0 > 0 there exists %i ∈ (0, 1) (which also depend on F ) so that for each x ∈ U and each v ∈ TxU

(1) each domain of Fi : Ui → Vi is δi-nice. (2) If U i does not contain a critical point of F and U i is compactly contained in x i Fi(x) Vi , then Fi(x) || DF (x)v|| i τ ||v|| i . i V > i Vx Fi(x) i (3) If Ux contains a critical point of F and d(x, Crit(F )) > κ0, then

|| DFi(x)v|| i %i||v||Vi . VF (x) > x i i (4) In fact, if x ∈ U , v ∈ TxU and d(x, Crit(F )) > κ0, then

|| DF (x)v|| i−1 % ||v|| i . i V > i Vx Fi(x) i i (5) If y ∈ V and w ∈ TyV , then

||w|| i−1 (1/%i)||w|| 0 Vy 6 Vy i i (we will apply this assertion for the case when x ∈ U and y = Fi(x) ∈ V ).

Proof. The niceness for Fi in Part (1) follows inductively from the niceness of Fi−1: since conformal maps preserve moduli and since one pulls back at most N = # Crit(F ) times by a non-conformal map, see [KvS, Lemma 8.3]. Part (2) follows since F : Ux → VF (x) is an isometry w.r.t. to the Poincar´emetric and the embedding Ux → Vx is a contraction (by a factor which only depends on δ). Part (3) holds because DFi only vanishes at critical points of F . THE DYNAMICS OF COMPLEX BOX MAPPINGS 81

i To see Part (4), we need to consider two subcases. (i) Fi−1(x) ∈ V and (ii) Fi−1(x) ∈ i−1 i V \V . If (i) holds then Fi(x) = Fi−1(x) and so by Part (2) we have

|| DF (x)v|| i−1 = || DF (x)v|| i−1 %||v|| i . i V i−1 V > Vx Fi(x) Fi−1(x) i−1 i t On the other hand, if (ii) holds, i.e. if Fi−1(x) ∈ V \V , then Fi is of the form Fi−1 where i−1 t > 2 and where at most one of the iterates is through a critical domain of Fi−1 : U → Vi−1 and the other iterates pass through diffeomorphic domains. Therefore again the assertion from Part (4) follows from Parts (1) and (2). Part (5) holds because the diameter of each component of Vi is bounded from below (with a bound which depends on F and n) and because one has a lower bound for the i−1 i modulus of the component of V \V containing y. 

Proposition 14.3. For each n > 0 and κ0 > 0 there exist τ > 1 and C > 0 so that the i following holds. Assume that 0 6 i 6 n, x ∈ U , d(x, Crit(F )) > κ0 and y = Fi(x) and t i define t = t(y, x), i.e. so that Fi(x) = F (x). Then for each v ∈ TxU one has t || DF (x)v|| i Cτ · ||v|| i , i Vy > Vx and t || DF (x)v|| i τ · ||v|| i . i Vy > Vx i if Ux does not contain a critical point. t−1 j Proof. Define t(j) = #{x, . . . , F (x) ∈ U }, let % = %i be as in the previous lemma, and choose η > 0 so that τ(%/τ)η > τ 0 := (1 + τ)/2. Let j be the smallest integer < i so that t(j) < ηt(j − 1). i−1 t(i−1)−1 Case 1: j does not exist. Then t(i − 1) > η t; write Fi(x) = Fi−1 ◦ Fi−1(x). The first of these iterates is through the central branch of Fi−1 and, since Fi is the first return back to Vi, the remaining iterates are non-critical branches. It follows as in the proof of Parts (4) and (1) of the previous lemma that t(i−1)−1 || DF (x)v|| i−1 % · τ · ||v|| i . i V > Vx Fi(x) i−1 Since t(i − 1) > η t, this implies t || DF (x)v|| i Cτ · ||v|| i , i Vy > Vx t by redefining τ. Let t0 be so large that Cτ > 1 for t > t0. Now takeτ ˆ > 1 so that t t t Cτi > τˆ for t > t0 and so that moreover τi > τˆ for all t 6 t0 (here τi corresponds to Part (1) of the previous lemma applied to Fi). By combining the first inequality and Part (1) of the previous lemma the second inequality follows by renamingτ ˆ as τ. j−1 Case 2: j does exist. Then t(j − 1) > η t and t(j) 6 η · t(j − 1). In that case iterate j−1 j j x by Fj−1 until an iterate enters V \V , and then iterate Fj−1 until we again hit V . This may repeat several times, but the net expansion we obtain is t(j) t(j−1)−t(j) || DF (x)v|| i−1 % τ · ||v|| i . i V > Vx Fi(x) THE DYNAMICS OF COMPLEX BOX MAPPINGS 82

Since τ(%/τ)η > τ 0 the first inequality follows. The 2nd inequality then holds as in Case 1.  The following lemma is straight-forward.

Lemma 14.4. There exists %0 > 0 so that if x ∈ V, v ∈ TxV, then

||v||V > %0||v||.

If d(x, ∂V) > κ, then there exists %1(κ) > 0 so that

%1||v||V 6 ||v||.  Let us now prove Theorem 14.1.

Proof of Theorem 14.1. We will decompose F k, by first mapping it via iterates of, suc- cessively, F and then of F ,...,F going down to closer and closer to the critical m0 m1 mi0 points of F (where m < m < ··· < m ) and then by maps F ,...,F (where 0 1 i0 mi0+1 mi1 k mi0 > mi0+1 > ··· > mi1 ) up to when the point F (x) has been reached. As we will iterate x not only by F but also by the maps Fm, we need to be sure we stop before we reach the k-th iterate of F . For this reason we define t = t(x, m, n) for x ∈ Vm n t so that Fm = F near x. This will allow us to make sure that if xi is an iterate of F of time < k, then we will only iterate this point with further iterates of F up to time < k. m0 m0+1 So let first describe m0 < ··· < mi0 . Let m0 > 0 be so that x0 := x ∈ V \V . i m0 m0+1 Then let n0 > 0 be maximal so that Fm0 (x0) ∈ V \V for all 0 6 i < n0 and so that t(x0, 0) := t(x0, m0, n0) 6 k. If n0 = 0, then define i0 = 0. If n0 > 0, then n0 t(x0,0) m0+1 m1 m1+1 let x1 := Fm0 (x0) = F (x0) ∈ V , let m1 be so that x1 ∈ V \V and n1 let n1 > 0 be maximal so that Fm1 (x1) ∈ Vm1 and so that t(x0, 1) := t(x1, m1, n1) + n1 t(x0,1) t(x0, 0) 6 k. Define x2 := Fm1 (x1) = F (x0) ∈ Vm1 . If n1 > 0 we continue and define inductively in a similar way xi, mi, ni, t(x0, i) for i = 0, . . . , i0 − 1 and xi0 where i0 mi0 is chosen maximal, i.e. xi0 ∈ V , but there is no F -iterate of x0 (up to k of F ) which mi0 +1 ni0 mi0 enters V . In this case let ni0 > 0 be maximal so that F (xi0 ) ∈ V and so that ni0 t(x ,i ) m t(x , i ) := t(x , m , n ) + t(x , i − 1) k. Define x = F (x ) = F 0 0 ∈ V i0 . 0 0 i0 i0 i0 0 0 6 i0+1 mi0 i0 mi mi0 Note that xi ∈ V for i = 0, . . . , i0 but that also xi0+1 ∈ V . Moreover, 0 6 m0 < m1 < ··· < mi0 and n0, . . . , ni0−1 > 0. Next we will define mi0+1, . . . , mi1 . Define mi0+1 maximal so that that t(x , m , 1)+t(x , i ) k, i.e. so that the iterate F (x ) i0+1 i0+1 0 0 6 mi0+1 i0+1 mi0+1 mi0 is before time k. Note that mi0+1 < mi0 and so xi0+1 ∈ V ⊂ V . Then take ni0+1 to be maximal so that t(x0, i0 + 1) := t(xi0+1, mi0+1, ni0+1) + t(x0, i0) 6 k and ni0+1 t(x ,i +1) m define x = F (x ) = F 0 0 (x ) ∈ V i0+1 . Next take m maximal so that i0+2 mi0+1 i0 0 i0+2 t(xi0+1, mi0+1, 1) + t(x0, i0 + 1) 6 k and define ni0+2 to be maximal so that t(x0, i0 + 2) := j(xi0+2, mi0+2, ni0+2)+t(x0, i0 +1) 6 k. Similarly, define mi, ni, t(x0, i) for i = i0 +1, . . . , i1 where i1 is maximal so that ‘time runs out’ (so further iterates would consider more than k iterates under F of x). THE DYNAMICS OF COMPLEX BOX MAPPINGS 83

Note that for i = i0 + 1, . . . , i1, the maximality of mi implies that

j mi+1 (14.1) ni > 1 =⇒ Fmi (xi) ∈/ V for j = 0, . . . , ni − 1 because otherwise one could iterate with Fmi+1 (before reaching time k). Let us now discuss expansion in terms of some Poincar´emetrics along this orbit. For simplicity choose τ > 1 and % > 0 so that these bound from below the constants τi,%i, i = 0, . . . , ν from Lemma 14.2 and so that the conclusion of Proposition 14.3 holds. For j i = 0, . . . , i0 and 0 6 j < ni, by definition Fmi (xi) is not contained in a critical domain of Vmi and so by Proposition 14.3 we get

j j t(Fmi (xi),xi) || DFmi (Fm (xi))v||V j+1 τ · ||v||V j+1 i F (x ) > F (x ) mi i mi i where τ > 1. This gives for i = 0, . . . , i0,

ni t(xi+1,xi) || DFm (xi)v|| mi τ · ||v|| mi . i Vxi+1 > Vxi

ni0−1 n m Note that x = F ◦ · · · ◦ F 0 (x ) ∈ V i0 and that the above statement gives i0 m0−1 m0 0

ni −1 0 n0 m t(xi0 ,x0) m || DFmi −1 ··· DFm (x0)(v)|| i0−1 τ · ||v|| 0 . 0 0 Vx > Vx0 i0

If ni0 = 0, then xi0+1 = xi0 and we obtain from Part (5) of Lemma 14.2 that n i0 n0 t(xi ,x0) || DF ··· DF (x )(v)|| 0 τ 0 · % · ||v|| m0 . mi0 m0 0 Vx > Vx i0+1 0

m mi −1 If n 1, then note that x ∈ V i0 and so for all v ∈ T U 0 i0 > i0+1 xi0 xi0 n i0 t(xi ,x0) || DF (x )(v)|| m −1 τ 0 · % · ||v|| mi −1 mi0 i0 i0 > 0 Vx Vxi i0+1 0 where we use for the first n − 1 iterates of F Proposition 14.3 and for the last iterate 0 mi0 Part (4) of Lemma 14.2. So in all cases we have

n i0 n0 t(xi +1,x0) 2 || DF ··· DF (x )(v)|| 0 τ 0 · % · ||v|| m0 . mi0 m0 0 Vx > Ux i0+1 0

Using the same argument for i = i0 + 1, . . . , i1 we thus obtain

ni k k 1 n0 t(F (x0),x0) 2 ν m || DF (x0)v|| 0 = || DFmi ··· DF (x0)(v)|| 0 τ · (% /τ) · ||v|| 0 . V k 1 m0 Vx > Ux F (x0) i0+1 0

k Using d(F (x), ∂V) > κ and Lemma 14.4 we can transfer this statement to one in terms of the Euclidean norm on C and obtain k k t(F (x0),x0) || DF (x0)v|| > C · τ · ||v|| 2 ν where C = (% /τ) (%1%0) and where ν is fixed (as in the statement of the theorem).  THE DYNAMICS OF COMPLEX BOX MAPPINGS 84

References [AIM] Kari Astala, Tadeusz Iwaniec, Gaven Martin, Elliptic partial differential equations and quasicon- formal mappings in the plane. Princeton Mathematical Series, 48. Princeton University Press, Princeton, NJ, 2009. xviii+677 pp. [AKLS] Artur Avila, Jeremy Kahn, Mikhail Lyubich, Weixiao Shen, Combinatorial rigidity for unicritical polynomials. Ann. of Math., 170 (2009), no. 2, 783–797. [AL1] Artur Avila, Mikhail Lyubich, Hausdorff dimension and conformal measures of Feigenbaum Julia sets. J. Amer. Math. Soc. 21 (2008), 305–363 . [AL2] Artur Avila, Mikhail Lyubich, The full renormalization horseshoe for unimodal maps of higherde- gree: Exponential contraction along hybrid classes. Publ. Math. IHES,´ 114 (2011), 171–223. [AL3] Artur Avila, Mikhail Lyubich, Lebesgue measure of Feigenbaum Julia sets. https://arxiv.org/ abs/1504.02986. [ALdM] Artur Avila, Mikhail Lyubich, Welington de Melo, Regular or stochastic dynamics in real analytic families of unimodal maps. Invent. Math. 154 (2003), 451–550. [ALS] Artur Avila, Mikhail Lyubich, Weixiao Shen, Parapuzzle of the Multibrot set and typical dynamics of unimodal maps. J. Eur. Math. Soc. 13 (2011), 27–56. [BL] Alexander Blokh, Mikhail Lyubich, of maps of the interval. Dynamical systems and ergodic theory. (Warsaw, 1986), 427–442, Banach Center Publ., 23, PWN, Warsaw, 1989. [BM] Mario Bonk, Daniel Meyer, Expanding Thurston maps, volume 225 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2017. [BSS] Henk Bruin, Weixiao Shen, Sebastian van Strien, Existence of unique SRB-measures is typical for real unicritical polynomial families. Ann. Sci. l’Ecole´ Norm. Sup. (4) 39 (2006), no. 3, 381–414. [BvS] Henk Bruin, Sebastian van Strien, Monotonicity of entropy for real multimodal maps. J. Amer. Math. Soc. 28 (2015), no. 1, 1–61. [Cl] Trevor Clark, Regular or stochastic dynamics in families of higher-degree unimodal maps. Ergodic Theory Dyn. Sys. 34 (2014), no. 5, 1538–1566. [CvS] Trevor Clark, Sebastian van Strien, Quasisymmetric rigidity in one-dimensional dynamics. https://arxiv.org/abs/1805.09284. [CvS2] Trevor Clark, Sebastian van Strien, Conjugacy classes of real analytic one-dimensional maps are analytic connected manifolds, manuscript. [CvST] Trevor Clark, Sebastian van Strien, Sofia Trejo, Complex Bounds for Real Maps. Comm. Math. Phys. 355 (2017), 1001–1119. [CT] Trevor Clark, Sofia Trejo, The boundary of chaos for interval mappings. Proc. Lond. Math. Soc. (3) 121 (2020), no. 6, 1427–1467. [dMP] Laura DeMarco, Kevin Pilgrim, The classification of polynomial basins of infinity. Ann. Sci. Ec.´ Norm. Sup´er.(4) 50 (2017), no. 4, 799–877. [Dob] Neil Dobbs, Nice sets and invariant densities in complex dynamics. Math. Proc. Camb. Phil. Soc. (1) 150 (2011), 157–165. [Do1] Adrien Douady, Descriptions of compact sets in C. In: Topological methods in modern mathe- matics (Stony Brook, NY, 1991), Publish or Perish, Houston, TX, 1993, 429–465. [Do2] Adrien Douady, Topological entropy of unimodal maps: monotonicity for quadratic polynomials, Real and complex dynamical systems (Hillerød, 1993) NATO Adv. Sci. Inst. Ser. C Math. Phys. Sci., vol. 464, Kluwer Acad. Publ., Dordrecht (1995), 65–87. [DH1] Adrien Douady, John Hubbard, On the dynamics of polynomial-like mappings. Ann. Sci. Ecole.´ Norm. Sup´er.4e series 18 (1985), 287–343. [DH2] Adrien Douady, John Hubbard, A proof of Thurston’s topological characterization of rational functions. Acta Math., 171 (1993), no. 2, 263–297. [Dr] Kostiantyn Drach, Rigidity of rational maps: an axiomatic approach, manuscript in preparation. THE DYNAMICS OF COMPLEX BOX MAPPINGS 85

[DMRS] Kostiantyn Drach, Yauhen Mikulich, Johannes R¨uckert, Dierk Schleicher, A combinatorial clas- sification of postcritically fixed Newton maps. Ergod. Theor. Dyn. Syst. 39 (2019), no. 11, 2983– 3014. [DLSS] Kostiantyn Drach, Russell Lodge, Dierk Schleicher, Maik Sowinski, Puzzles and the Fatou– Shishikura injection for rational Newton maps, Trans. Amer. Math. Soc. 374 (2021), no. 4, 2753–2784. [DS] Kostiantyn Drach, Dierk Schleicher, Rigidity of Newton dynamics, submitted, https://arxiv. org/abs/1812.11919. [GS] Jacek Graczyk, Grzegorz Swiatek,´ Generic hyperbolicity in the logistic family. Ann. Math. 146 (1997), 1–52. [GJ] John Guckenheimer, Stewart Johnson, Distortion of S-unimodal maps, Ann. of Math. 132 (1990), 71–130. [HK] Juha Heinonen, Pekka Koskela, Definitions of quasiconformality. Invent. Math. 120 (1995), 61– 79. [Hub1] John Hubbard, Local connectivity of Julia sets and bifurcation loci: three theorems of J.-C. Yoccoz. In: Topological Methods in Modern Mathematics. Publish or Perish, Houston (1992), 467–511. [Hub2] John Hubbard, Teichm¨uller theory and applications to geometry, topology, and dynamics, volume 1. Matrix Editions, Ithaca, 2006. [Hub3] John Hubbard, Teichm¨uller theory and applications to geometry, topology, and dynamics, volume 2. Matrix Editions, Ithaca, 2016. [Jak] Michael Jakobson, Absolutely continuous invariant measures for one-parameter families of one- dimensional maps. Commun. Math. Phys. 81 (1981), 39–88. [JS] Michael Jakobson, Grzegorz Swi¸atek,´ Metric properties of non-renormalizable S-unimodal maps. Part I: Induced expansion and invariant measures. Ergod. Theor. Dyn. Syst. 14 (1994), 721–755. [Ji] Yunping Jiang, Renormalization and geometry in one-dimensional and complex dynamics. Ad- vanced Series in Nonlinear Dynamics, 10. World Scientific Publishing Co., Inc., River Edge, NJ, 1996. xvi+309. [Ka] Jeremy Kahn, Holomorphic Removability of Julia Sets. https://arxiv.org/abs/math/9812164. [KL1] Jeremy Kahn, Mikhail Lyubich, The quasi-additivity law in conformal geometry. Ann. Math. 169 (2009), no. 2, 561–593. [KL2] Jeremy Kahn, Mikhail Lyubich, Local connectivity of Julia sets for unicritical polynomials. Ann. Math. 170 (2009), no. 1, 413–426. [Ko1] Oleg Kozlovski, Axiom A maps are dense in the space of unimodal maps in the Ck topology. Ann. Math. 157 (2003), no. 1, 1–43. [Ko2] Oleg Kozlovski, On the structure of isentropes of real polynomials. J. Lond. Math. Soc. (2) 100 (2019), no. 1, 159–182. [KSS1] Oleg Kozlovski, Weixiao Shen, Sebastian van Strien, Rigidity for real polynomials. Ann. Math., 165 (2007), 749–841. [KSS2] Oleg Kozlovski, Weixiao Shen, Sebastian van Strien, Density of hyperbolicity in dimension one. Ann. Math. 166 (2007), 145–182. [KvS] Oleg Kozlovski, Sebastian van Strien, Local connectivity and quasi-conformal rigidity of non- renormalizable polynomials. Proc. London Math. Soc. (3) 99 (2009), 275–296. [LvS1] Genadi Levin, Sebastian van Strien, Local connectivity of the Julia set of real polynomials. Ann. Math. 147 (1998), 471–541. [LvS2] Genadi Levin, Sebastian van Strien, Bounds for maps of an interval with one critical point of inflection type. II. Invent. Math. 141 (2000), 399–465. [LvS3] Genadi Levin, Sebastian van Strien, Total disconnectedness of Julia sets and absence of invari- ant linefields for real polynomials. in G´eom´etrie complexe et syst`emesdynamiques—Colloque en l’honneur d’Adrien Douady Orsay, 1995, Ast´erisque,no. 261 (2000), pp. 161–172. THE DYNAMICS OF COMPLEX BOX MAPPINGS 86

[Lyu1] Mikhail Lyubich, On the Lebesgue measure of the Julia set of a quadratic polynomial. https: //arxiv.org/abs/math/9201285 [Lyu2] Mikhail Lyubich, Dynamics of quadratic polynomials. I, II. Acta Math. 178 (1997), no. 2, 185– 297. [Lyu3] Mikhail Lyubich, Feigenbaum-Coullet-Tresser universality and Milnor’s Hairiness Conjecture. Ann. Math. 149 (1999), 319–420. [Lyu4] Mikhail Lyubich, Dynamics of quadratic polynomials III. In: G´eom´etriecomplexe et syst`emes dynamiques—Colloque en l’honneur d’Adrien Douady Orsay, 1995, Ast´erisque, 261 (2000), pp. 173–200. [Lyu5] Mikhail Lyubich, Almost every real quadratic map is either regular or stochastic. Ann. Math. 156 (2002), 1–78. [Lyu6] Mikhail Lyubich, Teichm¨uller Space of Fibonacci Maps. In: Pacifico M., Guarino P. (eds) New Trends in One-Dimensional Dynamics. Springer Proceedings in Mathematics & Statistics, vol 285. Springer, Cham (2019) 221–237. [Lyu7] Mikhail Lyubich, Conformal Geometry and Dynamics of Quadratic Polynomials, vol I–II, book in preparation, http://www.math.stonybrook.edu/~mlyubich/book.pdf. [LY] Mikhail Lyubich, Michael Yampolsky, Dynamics of quadratic polynomials: Complex bounds for real maps. Ann. Inst. Fourier (Grenoble) 47 (1997), no. 4, 1219–1255. [LM] Mikhail Lyubich, John Milnor, The Fibonacci unimodal map. J. Amer. Math. Soc. 6 (1993), no. 2, 425–457. [Man] Richardo Ma˜n´e, On a theorem of Fatou. Bol. Soc. Brasil. Mat. (N.S.), 24 (1993), no. 1, 1–11. [Mar] Marco Martens, Distortion Results and Invariant Cantor sets for Unimodal maps. Ergod. Th. Dynam. Sys. 14 (1994), 331–349. [McM] Curtis McMullen, Complex dynamics and renormalization. Annals of Mathematics Studies, 135, Princeton University Press, Princeton, NJ, 1994. [McM2] Curtis McMullen, Renormalization and 3-manifolds which fiber over the circle. Annals of Math- ematics Studies, 142, Princeton University Press, Princeton, NJ, 1996. [dMvS] Welington de Melo, Sebastian van Strien, One-dimensional dynamics. Springer-Verlag, New York, 1993. [Mil1] John Milnor, Local connectivity of Julia sets: expository lectures. The Mandelbrot set, theme and variations, 67–116, London Math. Soc. Lecture Note Ser., 274, Cambridge Univ. Press, Cambridge, 2000. [Mil2] John Milnor, Dynamics in one complex variable, third edition. Annals of Mathematics Studies, 160, Princeton University Press, Princeton, NJ, 2006. [Mil3] John Milnor, On Latt`esmaps. In: Hjorth, P., Petersen, C. L., Dynamics on the Riemann Sphere. A Bodil Branner Festschrift, Eur. Math. Soc., Z¨urich (2006) 9–43. [Mil4] John Milnor, Non locally-connected Julia sets constructed by iterated turning. Birthday lecture for ‘Douady 70’. Revised May 26, 2006, SUNY, 28 pp, https://www.math.stonybrook.edu/~jack/ tune-b.pdf [MTh] John Milnor, William Thurston, On iterated maps of the interval, Dynamical systems (College Park, MD, 1986), Lecture Notes in Math., vol. 1342, Springer, Berlin, 1988, pp. 465–563. [MTr] John Milnor, Charles Tresser, On entropy and monotonicity for real cubic maps. Comm. Math. Phys. 209 (2000), no. 1, 123–178. With an appendix by Adrien Douady and Pierrette Sentenac. [NvS] Tomasz Nowicki, Sebastian van Strien, Invariant measures exist under a summability condition for unimodal maps. Invent. Math. 105 (1991), no. 1, 123–136. [PR] Carsten Petersen, Pascale Roesch, Parabolic tools. J. Difference Eq. Appl., 16 (2010), no. 5-6, 715–738. [Prz] Feliks Przytycki, Interations of Holomorphic Collet-Eckmann Maps: Conformal and Invariant Measures. Appendix: On Non-Renormalizable Quadratic Polynomials. Trans. Amer. Math. Soc., 350 (1998), no. 2, 717–742. THE DYNAMICS OF COMPLEX BOX MAPPINGS 87

[PRL1] Feliks Przytycki, Juan Rivera-Letelier, Statistical properties of topological Collet–Eckmann maps. Ann. Sci. Ecole´ Norm. Sup. 40 (2007), no. 1, 135–178. [PRL2] Feliks Przytycki, Juan Rivera-Letelier, Nice inducing schemes and the thermodynamics of rational maps. Commun. Math. Phys. 301 (2011), 661–707. [PZ] Feliks Przytycki, Anna Zdunik, On Hausdorff dimension of polynomial not totally disconnected Julia sets. https://arxiv.org/abs/2003.12612. [RvS] Lasse Rempe-Gillen, Sebastian van Strien, Density of hyperbolicity for classes of real transcen- dental entire functions and circle maps. Duke Math. J. 164 (2015), no. 6, 1079–1137. [Ri] Seppo Rickman, Removability theorems for quasiconformal mappings. Ann. Ac. Scient. Fenn. 449 (1969), 1–8. [R-L] Juan Rivera-Letelier, A connecting lemma for rational maps satisfying a no-growth condition. Ergodic Theory and Dynamical Systems, 27 (2007), no. 2, 595–636. [RLS] Juan Rivera-Letelier, Weixiao Shen, Statistical properties of one-dimensional maps under weak hyperbolicity assumptions. Ann. Sci. Ecole´ Norm. Sup. 47 (2014), no. 6, 1027–1083. [Roe] Pascale Roesch, On local connectivity for the Julia set of rational maps: Newton’s famous example. Ann. Math. 168 (2008), 1–48. [RY] Pascale Roesch, Yongcheng Yin, The boundary of bounded polynomial Fatou components. C. R. Math. Acad. Sci. Paris 346 (2008), no. 15-16, 877–880 (https://arxiv.org/abs/0909.4598v2). [Sch1] Dierk Schleicher, On fibers and local connectivity of compact sets in C. Stony Brook IMS preprint #1998/12. [Sch2] Dierk Schleicher, On fibers and local connectivity of Mandelbrot and Multibrot sets. In: M. Lapidus, M. van Frankenhuysen (eds): Fractal Geometry and Applications: A Jubilee of . Proceedings of Symposia in Pure Mathematics 72, American Mathematical Society (2004), 477–507. [Sch3] Dierk Schleicher, appendix to William Thurston, On the geometry and dynamics of iterated rational maps. Edited by Dierk Schleicher and Nikita Selinger. In D. Schleicher, editor, Complex dynamics: families and friends, A K Peters, Wellesley, MA, 2009; 3–137. [She1] Weixiao Shen, On the measurable dynamics of real rational functions. Ergod. Theory Dyn. Syst. 23 (2003), 957–983. [She2] Weixiao Shen, On the metric properties of multimodal interval maps and C2 density of Axiom A. Invent. Math. 156 (2004), 310–401. [Shi] Mitsuhiro Shishikura, Unpublished. [ST] Mitsuhiro Shishikura, Tan Lei, An alternative proof of Ma˜n´e’stheorem on non-expanding Julia sets. The Mandelbrot set, theme and variations, 265–279, London Math. Soc. Lecture Note Ser., 274, Cambridge Univ. Press, Cambridge, 2000. [Sm1] Daniel Smania, Complex bounds for multimodal maps: bounded combinatorics. Nonlinearity 14 (2001), no. 5, 1311–1330. [Sm2] Daniel Smania, Phase space universality for multimodal maps. Bull. Braz. Math. Soc. 36 (2005), no. 2, 225–274. [Sm3] Daniel Smania, Puzzle geometry and rigidity: the Fibonacci cycle is hyperbolic. J. Amer. Math. Soc. 20 (2007), no. 3, 629–673. [SV] Sebastian van Strien, Edson Vargas, Real bounds, ergodicity and negative Schwarzian for multi- modal maps. J. Amer. Math. Soc. 17 (2004), no. 4, 749–782. [Su] Dennis Sullivan, Bounds, quadratic differentials and renormalization conjectures, MS Centennial Publications, 2: Mathematics into Twenty-first Century, 1992. [Ts] Masato Tsujii, A simple proof for monotonicity of entropy in the quadratic family. Ergod. Theory Dyn. Syst. 20 (2000), no. 3, 925–933. [QY] Weiyuan Qiu, Yongcheng Yin, Proof of the Branner–Hubbard conjecture on Cantor Julia sets. Sci. China Ser. A 52 (2009), 45–65. THE DYNAMICS OF COMPLEX BOX MAPPINGS 88

[QWY] Weiyuan Qiu, Xiaoguang Wang, Yongcheng Yin, Dynamics of McMullen maps. Adv. Math. 229 (2012), 2525–2577. [WYZ] Xiaoguang Wang, Yongcheng Yin, Jinsong Zeng, Dynamics of Newton maps. https://arxiv. org/abs/1805.11478. [Ya] Jonguk Yang, Local connectivity of polynomial Julia sets at bounded type Siegel boundaries. https: //arxiv.org/abs/2010.14003. [YZ] Yongcheng Yin, Yu Zhai, No invariant line fields on Cantor Julia sets. Forum Math. 22 (2010), 75–94. [Zh] Yu Zhai, On the ergodicity of conformal measures for rational maps with totally disconnected Julia sets. Proc. Am. Math. Soc. 140 (2012), no. 10, 3453–3462.

Open University, School of Mathematics and Statistics, Milton Keynes MK7 6AA, UK Email address: [email protected]

Aix–Marseille Universite,´ Institut de Mathematiques´ de Marseille, 163 Avenue de Luminy, 13009 Marseille, France Email address: [email protected]

University of Warwick, Mathematics Institute, Zeeman Building, Coventry CV4 7AL, UK Email address: [email protected]

Imperial College London, Department of Mathematics, 180 Queen’s Gate, London SW7 2AZ, UK Email address: [email protected]